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 31 毫秒
1.
Meneely KM  Lamb AL 《Biochemistry》2007,46(42):11930-11937
Pyoverdin is the hydroxamate siderophore produced by the opportunistic pathogen Pseudomonas aeruginosa under the iron-limiting conditions of the human host. This siderophore includes derivatives of ornithine in the peptide backbone that serve as iron chelators. PvdA is the ornithine hydroxylase, which performs the first enzymatic step in preparation of these derivatives. PvdA requires both flavin adenine dinucleotide (FAD) and nicotinamide adenine dinucleotide phosphate (NADPH) for activity; it was found to be a soluble monomer most active at pH 8.0. The enzyme demonstrated Michaelis-Menten kinetics in an NADPH oxidation assay, but a hydroxylation assay indicated substrate inhibition at high ornithine concentration. PvdA is highly specific for both substrate and coenzyme, and lysine was shown to be a nonsubstrate effector and mixed inhibitor of the enzyme with respect to ornithine. Chloride is a mixed inhibitor of PvdA with respect to ornithine but a competitive inhibitor with respect to NADPH, and a bulky mercurial compound (p-chloromercuribenzoate) is a mixed inhibitor with respect to ornithine. Steady-state experiments indicate that PvdA/FAD forms a ternary complex with NADPH and ornithine for catalysis. PvdA in the absence of ornithine shows slow substrate-independent flavin reduction by NADPH. Biochemical comparison of PvdA to p-hydroxybenzoate hydroxylase (PHBH, from Pseudomonas fluorescens) and flavin-containing monooxygenases (FMOs, from Schizosaccharomyces pombe and hog liver microsomes) leads to the hypothesis that PvdA catalysis proceeds by a novel reaction mechanism.  相似文献   

2.
In the malarial parasite Plasmodium falciparum, a multifunctional phosphoethanolamine methyltransferase (PfPMT) catalyzes the methylation of phosphoethanolamine (pEA) to phosphocholine for membrane biogenesis. This pathway is also found in plant and nematodes, but PMT from these organisms use multiple methyltransferase domains for the S-adenosylmethionine (AdoMet) reactions. Because PfPMT is essential for normal growth and survival of Plasmodium and is not found in humans, it is an antiparasitic target. Here we describe the 1.55 Å resolution crystal structure of PfPMT in complex with AdoMet by single-wavelength anomalous dispersion phasing. In addition, 1.19–1.52 Å resolution structures of PfPMT with pEA (substrate), phosphocholine (product), sinefungin (inhibitor), and both pEA and S-adenosylhomocysteine bound were determined. These structures suggest that domain rearrangements occur upon ligand binding and provide insight on active site architecture defining the AdoMet and phosphobase binding sites. Functional characterization of 27 site-directed mutants identifies critical active site residues and suggests that Tyr-19 and His-132 form a catalytic dyad. Kinetic analysis, isothermal titration calorimetry, and protein crystallography of the Y19F and H132A mutants suggest a reaction mechanism for the PMT. Not only are Tyr-19 and His-132 required for phosphobase methylation, but they also form a “catalytic” latch that locks ligands in the active site and orders the site for catalysis. This study provides the first insight on this antiparasitic target enzyme essential for survival of the malaria parasite; however, further studies of the multidomain PMT from plants and nematodes are needed to understand the evolutionary division of metabolic function in the phosphobase pathway of these organisms.  相似文献   

3.
The polyisoprenoid compound undecaprenyl phosphate is required for biosynthesis of cell wall peptidoglycans in Gram-positive bacteria, including pathogenic Enterococcus, Streptococcus, and Staphylococcus spp. In these organisms, the mevalonate pathway is used to produce the precursor isoprenoid, isopentenyl 5-diphosphate. Mevalonate diphosphate decarboxylase (MDD) catalyzes formation of isopentenyl 5-diphosphate in an ATP-dependent irreversible reaction and is therefore an attractive target for inhibitor development that could lead to new antimicrobial agents. To facilitate exploration of this possibility, we report the crystal structure of Staphylococcus epidermidis MDD (1.85 Å resolution) and, to the best of our knowledge, the first structures of liganded MDD. These structures include MDD bound to the mevalonate 5-diphosphate analogs diphosphoglycolyl proline (2.05 Å resolution) and 6-fluoromevalonate diphosphate (FMVAPP; 2.2 Å resolution). Comparison of these structures provides a physical basis for the significant differences in Ki values observed for these inhibitors. Inspection of enzyme/inhibitor structures identified the side chain of invariant Ser192 as making potential contributions to catalysis. Significantly, Ser → Ala substitution of this side chain decreases kcat by ∼103-fold, even though binding interactions between FMVAPP and this mutant are similar to those observed with wild type MDD, as judged by the 2.1 Å cocrystal structure of S192A with FMVAPP. Comparison of microbial MDD structures with those of mammalian counterparts reveals potential targets at the active site periphery that may be exploited to selectively target the microbial enzymes. These studies provide a structural basis for previous observations regarding the MDD mechanism and inform future work toward rational inhibitor design.  相似文献   

4.
The β-hydroxyacid dehydrogenases form a large family of ubiquitous enzymes that catalyze oxidation of various β-hydroxy acid substrates to corresponding semialdehydes. Several known enzymes include β-hydroxyisobutyrate dehydrogenase, 6-phosphogluconate dehydrogenase, 2-(hydroxymethyl)glutarate dehydrogenase, and phenylserine dehydrogenase, but the vast majority of β-hydroxyacid dehydrogenases remain uncharacterized. Here, we demonstrate that the predicted β-hydroxyisobutyrate dehydrogenase PA0743 from Pseudomonas aeruginosa catalyzes an NAD+-dependent oxidation of l-serine and methyl-l-serine but exhibits low activity against β-hydroxyisobutyrate. Two crystal structures of PA0743 were solved at 2.2–2.3-Å resolution and revealed an N-terminal Rossmann fold domain connected by a long α-helix to the C-terminal all-α domain. The PA0743 apostructure showed the presence of additional density modeled as HEPES bound in the interdomain cleft close to the predicted catalytic Lys-171, revealing the molecular details of the PA0743 substrate-binding site. The structure of the PA0743-NAD+ complex demonstrated that the opposite side of the enzyme active site accommodates the cofactor, which is also bound near Lys-171. Site-directed mutagenesis of PA0743 emphasized the critical role of four amino acid residues in catalysis including the primary catalytic residue Lys-171. Our results provide further insight into the molecular mechanisms of substrate selectivity and activity of β-hydroxyacid dehydrogenases.  相似文献   

5.
N-Hydroxylating monooxygenases are involved in the biosynthesis of iron-chelating hydroxamate-containing siderophores that play a role in microbial virulence. These flavoenzymes catalyze the NADPH- and oxygen-dependent hydroxylation of amines such as those found on the side chains of lysine and ornithine. In this work we report the biochemical and structural characterization of Nocardia farcinica Lys monooxygenase (NbtG), which has similar biochemical properties to mycobacterial homologs. NbtG is also active on d-Lys, although it binds l-Lys with a higher affinity. Differently from the ornithine monooxygenases PvdA, SidA, and KtzI, NbtG can use both NADH and NADPH and is highly uncoupled, producing more superoxide and hydrogen peroxide than hydroxylated Lys. The crystal structure of NbtG solved at 2.4 Å resolution revealed an unexpected protein conformation with a 30° rotation of the NAD(P)H domain with respect to the flavin adenine dinucleotide (FAD) domain that precludes binding of the nicotinamide cofactor. This “occluded” structure may explain the biochemical properties of NbtG, specifically with regard to the substantial uncoupling and limited stabilization of the C4a-hydroperoxyflavin intermediate. Biological implications of these findings are discussed.  相似文献   

6.
Enzymes of the de novo purine biosynthetic pathway have been identified as essential for the growth and survival of Mycobacterium tuberculosis and thus have potential for the development of anti-tuberculosis drugs. The final two steps of this pathway are carried out by the bifunctional enzyme 5-aminoimidazole-4-carboxamide ribonucleotide transformylase/inosine monophosphate cyclohydrolase (ATIC), also known as PurH. This enzyme has already been the target of anti-cancer drug development. We have determined the crystal structures of the M. tuberculosis ATIC (Rv0957) both with and without the substrate 5-aminoimidazole-4-carboxamide ribonucleotide, at resolutions of 2.5 and 2.2 Å, respectively. As for other ATIC enzymes, the protein is folded into two domains, the N-terminal domain (residues 1–212) containing the cyclohydrolase active site and the C-terminal domain (residues 222–523) containing the formyltransferase active site. An adventitiously bound nucleotide was found in the cyclohydrolase active site in both structures and was identified by NMR and mass spectral analysis as a novel 5-formyl derivative of an earlier intermediate in the biosynthetic pathway 4-carboxy-5-aminoimidazole ribonucleotide. This result and other studies suggest that this novel nucleotide is a cyclohydrolase inhibitor. The dimer formed by M. tuberculosis ATIC is different from those seen for human and avian ATICs, but it has a similar ∼50-Å separation of the two active sites of the bifunctional enzyme. Evidence in M. tuberculosis ATIC for reactivity of half-the-sites in the cyclohydrolase domains can be attributed to ligand-induced movements that propagate across the dimer interface and may be a common feature of ATIC enzymes.  相似文献   

7.
Abietadiene synthase from Abies grandis (AgAS) is a model system for diterpene synthase activity, catalyzing class I (ionization-initiated) and class II (protonation-initiated) cyclization reactions. Reported here is the crystal structure of AgAS at 2.3 Å resolution and molecular dynamics simulations of that structure with and without active site ligands. AgAS has three domains (α, β, and γ). The class I active site is within the C-terminal α domain, and the class II active site is between the N-terminal γ and β domains. The domain organization resembles that of monofunctional diterpene synthases and is consistent with proposed evolutionary origins of terpene synthases. Molecular dynamics simulations were carried out to determine the effect of substrate binding on enzymatic structure. Although such studies of the class I active site do lead to an enclosed substrate-Mg2+ complex similar to that observed in crystal structures of related plant enzymes, it does not enforce a single substrate conformation consistent with the known product stereochemistry. Simulations of the class II active site were more informative, with observation of a well ordered external loop migration. This “loop-in” conformation not only limits solvent access but also greatly increases the number of conformational states accessible to the substrate while destabilizing the nonproductive substrate conformation present in the “loop-out” conformation. Moreover, these conformational changes at the class II active site drive the substrate toward the proposed transition state. Docked substrate complexes were further assessed with regard to the effects of site-directed mutations on class I and II activities.  相似文献   

8.
Cytochrome P450SPα (CYP152B1) isolated from Sphingomonas paucimobilis is the first P450 to be classified as a H2O2-dependent P450. P450SPα hydroxylates fatty acids with high α-regioselectivity. Herein we report the crystal structure of P450SPα with palmitic acid as a substrate at a resolution of 1.65 Å. The structure revealed that the Cα of the bound palmitic acid in one of the alternative conformations is 4.5 Å from the heme iron. This conformation explains the highly selective α-hydroxylation of fatty acid observed in P450SPα. Mutations at the active site and the F–G loop of P450SPα did not impair its regioselectivity. The crystal structures of mutants (L78F and F288G) revealed that the location of the bound palmitic acid was essentially the same as that in the WT, although amino acids at the active site were replaced with the corresponding amino acids of cytochrome P450BSβ (CYP152A1), which shows β-regioselectivity. This implies that the high regioselectivity of P450SPα is caused by the orientation of the hydrophobic channel, which is more perpendicular to the heme plane than that of P450BSβ.  相似文献   

9.
3-Ketosteroid Δ4-(5α)-dehydrogenases (Δ4-(5α)-KSTDs) are enzymes that introduce a double bond between the C4 and C5 atoms of 3-keto-(5α)-steroids. Here we show that the ro05698 gene from Rhodococcus jostii RHA1 codes for a flavoprotein with Δ4-(5α)-KSTD activity. The 1.6 Å resolution crystal structure of the enzyme revealed three conserved residues (Tyr-319, Tyr-466, and Ser-468) in a pocket near the isoalloxazine ring system of the FAD co-factor. Site-directed mutagenesis of these residues confirmed that they are absolutely essential for catalytic activity. A crystal structure with bound product 4-androstene-3,17-dione showed that Ser-468 is in a position in which it can serve as the base abstracting the 4β-proton from the C4 atom of the substrate. Ser-468 is assisted by Tyr-319, which possibly is involved in shuttling the proton to the solvent. Tyr-466 is at hydrogen bonding distance to the C3 oxygen atom of the substrate and can stabilize the keto-enol intermediate occurring during the reaction. Finally, the FAD N5 atom is in a position to be able to abstract the 5α-hydrogen of the substrate as a hydride ion. These features fully explain the reaction catalyzed by Δ4-(5α)-KSTDs.  相似文献   

10.
The first step of the shikimate pathway for aromatic amino acid biosynthesis is catalyzed by 3-deoxy-d-arabino-heptulosonate 7-phosphate synthase (DAH7PS). Thermotoga maritima DAH7PS (TmaDAH7PS) is tetrameric, with monomer units comprised of a core catalytic (β/α)8 barrel and an N-terminal domain. This enzyme is inhibited strongly by tyrosine and to a lesser extent by the presence of phenylalanine. A truncated mutant of TmaDAH7PS lacking the N-terminal domain was catalytically more active and completely insensitive to tyrosine and phenylalanine, consistent with a role for this domain in allosteric inhibition. The structure of this protein was determined to 2.0 Å. In contrast to the wild-type enzyme, this enzyme is dimeric. Wild-type TmaDAH7PS was co-crystallized with tyrosine, and the structure of this complex was determined to a resolution of 2.35 Å. Tyrosine was found to bind at the interface between two regulatory N-terminal domains, formed from diagonally located monomers of the tetramer, revealing a major reorganization of the regulatory domain with respect to the barrel relative to unliganded enzyme. This significant conformational rearrangement observed in the crystal structures was also clearly evident from small angle X-ray scattering measurements recorded in the presence and absence of tyrosine. The closed conformation adopted by the protein on tyrosine binding impedes substrate entry into the neighboring barrel, revealing an unusual tyrosine-controlled gating mechanism for allosteric control of this enzyme.  相似文献   

11.
The biosynthesis of UDP-N-acetylmuramic acid (UDP-MurNAc) by reduction of UDP-N-acetylglucosamine-enolpyruvate (UDP-GlcNAc-EP) in an NADPH and FAD-dependent reaction in bacteria is one of the key steps in peptidoglycan biosynthesis catalyzed by UDP-N-acetylglucosamine-enolpyruvate reductase (MurB). Here, we present the crystal structure of Mycobacterium tuberculosis MurB (MtbMurB) with FAD as the prosthetic group at 2.0 Å resolution. There are six molecules in asymmetric unit in the form of dimers. Each protomer can be subdivided into three domains and the prosthetic group, FAD is bound in the active site between domain I and domain II. Comparison of MtbMurB structure with the structures of the Escherichia coli MurB (in complex with UDP-GlcNAc-EP) and Pseudomonas aeruginosa MurB (in complex with NADPH) showed all three structures share similar domain architecture and residues in the active site. The nicotinamide and the enol pyruvyl moieties are well aligned upon superimposition, both positioned in suitable position for hydride transfer to and from FAD. The comparison studies and MD simulations demonstrate that the two lobes of domain-III become more flexible. The substrates (NADPH and UDP-GlcNAc-EP) binding responsible for open conformation of MurB, suggesting that NADPH and UDP-GlcNAc-EP interactions are conformationally stable. Our findings provide a detail mechanism about the closed to open state by binding of NADPH and UDP-GlcNAc-EP induces the conformational changes of MurB structure that may trigger the MurB catalytic reaction.  相似文献   

12.
Cysteine (Cys) plays a major role in growth and survival of the human parasite Entamoeba histolytica. We report here the crystal structure of serine acetyltransferase (SAT) isoform 1, a cysteine biosynthetic pathway enzyme from E. histolytica (EhSAT1) at 1.77 Å, in complex with its substrate serine (Ser) at 1.59 Å and inhibitor Cys at 1.78 Å resolution. EhSAT1 exists as a trimer both in solution as well as in crystal structure, unlike hexamers formed by other known SATs. The difference in oligomeric state is due to the N-terminal region of the EhSAT1, which has very low sequence similarity to known structures, also differs in orientation and charge distribution. The Ser and Cys bind to the same site, confirming that Cys is a competitive inhibitor of Ser. The disordered C-terminal region and the loop near the active site are responsible for solvent-accessible acetyl-CoA binding site and, thus, lose inhibition to acetyl-CoA by the feedback inhibitor Cys. Docking and fluorescence studies show that EhSAT1 C-terminal-mimicking peptides can bind to O-acetyl serine sulfhydrylase (EhOASS), whereas native C-terminal peptide does not show any binding. To test further, C-terminal end of EhSAT1 was mutated and found that it inhibits EhOASS, confirming modified EhSAT1 can bind to EhOASS. The apparent inability of EhSAT1 to form a hexamer and differences in the C-terminal region are likely to be the major reasons for the lack of formation of the large cysteine synthase complex and loss of a complex regulatory mechanism in E. histolytica.  相似文献   

13.
The opium poppy (Papaver somniferum L.) is one of the oldest known medicinal plants. In the biosynthetic pathway for morphine and codeine, salutaridine is reduced to salutaridinol by salutaridine reductase (SalR; EC 1.1.1.248) using NADPH as coenzyme. Here, we report the atomic structure of SalR to a resolution of ∼1.9 Å in the presence of NADPH. The core structure is highly homologous to other members of the short chain dehydrogenase/reductase family. The major difference is that the nicotinamide moiety and the substrate-binding pocket are covered by a loop (residues 265–279), on top of which lies a large “flap”-like domain (residues 105–140). This configuration appears to be a combination of the two common structural themes found in other members of the short chain dehydrogenase/reductase family. Previous modeling studies suggested that substrate inhibition is due to mutually exclusive productive and nonproductive modes of substrate binding in the active site. This model was tested via site-directed mutagenesis, and a number of these mutations abrogated substrate inhibition. However, the atomic structure of SalR shows that these mutated residues are instead distributed over a wide area of the enzyme, and many are not in the active site. To explain how residues distal to the active site might affect catalysis, a model is presented whereby SalR may undergo significant conformational changes during catalytic turnover.  相似文献   

14.
The crystal structure of the full-length rat peroxisomal multifunctional enzyme, type 1 (rpMFE1), has been determined at 2.8 Å resolution. This enzyme has three catalytic activities and two active sites. The N-terminal part has the crotonase fold, which builds the active site for the Δ32-enoyl-CoA isomerase and the Δ2-enoyl-CoA hydratase-1 catalytic activities, and the C-terminal part has the (3S)-hydroxyacyl-CoA dehydrogenase fold and makes the (3S)-hydroxyacyl-CoA dehydrogenase active site. rpMFE1 is a multidomain protein having five domains (A–E). The crystal structure of full-length rpMFE1 shows a flexible arrangement of the A-domain with respect to the B–E-domains. Because of a hinge region near the end of the A-domain, two different positions of the A-domain were observed for the two protein molecules (A and B) of the asymmetric unit. In the most closed conformation, the mode of binding of CoA is stabilized by domains A and B (helix-10), as seen in other crotonase fold members. Domain B, although functionally belonging to the N-terminal part, is found tightly associated with the C-terminal part, i.e. fixed to the E-domain. The two active sites of rpMFE1 are ∼40 Å apart, separated by a tunnel, characterized by an excess of positively charged side chains. Comparison of the structures of rpMFE1 with the monofunctional crotonase and (3S)-hydroxyacyl-CoA dehydrogenase superfamily enzymes, as well as with the bacterial α2β2-fatty acid oxidation multienzyme complex, reveals that this tunnel could be important for substrate channeling, as observed earlier on the basis of the kinetics of rpMFE1 purified from rat liver.  相似文献   

15.
Biosynthesis of the mycobacterial cell wall relies on the activities of many enzymes, including several glycosyltransferases (GTs). The polymerizing galactofuranosyltransferase GlfT2 (Rv3808c) synthesizes the bulk of the galactan portion of the mycolyl-arabinogalactan complex, which is the largest component of the mycobacterial cell wall. We used x-ray crystallography to determine the 2.45-Å resolution crystal structure of GlfT2, revealing an unprecedented multidomain structure in which an N-terminal β-barrel domain and two primarily α-helical C-terminal domains flank a central GT-A domain. The kidney-shaped protomers assemble into a C4-symmetric homotetramer with an open central core and a surface containing exposed hydrophobic and positively charged residues likely involved with membrane binding. The structure of a 3.1-Å resolution complex of GlfT2 with UDP reveals a distinctive mode of nucleotide recognition. In addition, models for the binding of UDP-galactofuranose and acceptor substrates in combination with site-directed mutagenesis and kinetic studies suggest a mechanism that explains the unique ability of GlfT2 to generate alternating β-(1→5) and β-(1→6) glycosidic linkages using a single active site. The topology imposed by docking a tetrameric assembly onto a membrane bilayer also provides novel insights into aspects of processivity and chain length regulation in this and possibly other polymerizing GTs.  相似文献   

16.
17.
Pyridoxal 5′-phosphate (PLP)-dependent enzymes utilize the unique chemistry of a pyridine ring to carry out diverse reactions involving amino acids. Diaminopropionate (DAP) ammonia-lyase (DAPAL) is a prokaryotic PLP-dependent enzyme that catalyzes the degradation of d- and l-forms of DAP to pyruvate and ammonia. Here, we report the first crystal structure of DAPAL from Escherichia coli (EcDAPAL) in tetragonal and monoclinic forms at 2.0 and 2.2 Å resolutions, respectively. Structures of EcDAPAL soaked with substrates were also determined. EcDAPAL has a typical fold type II PLP-dependent enzyme topology consisting of a large and a small domain with the active site at the interface of the two domains. The enzyme is a homodimer with a unique biological interface not observed earlier. Structure of the enzyme in the tetragonal form had PLP bound at the active site, whereas the monoclinic structure was in the apo-form. Analysis of the apo and holo structures revealed that the region around the active site undergoes transition from a disordered to ordered state and assumes a conformation suitable for catalysis only upon PLP binding. A novel disulfide was found to occur near a channel that is likely to regulate entry of ligands to the active site. EcDAPAL soaked with dl-DAP revealed density at the active site appropriate for the reaction intermediate aminoacrylate, which is consistent with the observation that EcDAPAL has low activity under crystallization conditions. Based on the analysis of the structure and results of site-directed mutagenesis, a two-base mechanism of catalysis involving Asp120 and Lys77 is suggested.  相似文献   

18.
Fructose-1,6-bisphosphatase (FBPase), a key enzyme of gluconeogenesis and photosynthetic CO2 fixation, catalyzes the hydrolysis of fructose 1,6-bisphosphate (FBP) to produce fructose 6-phosphate, an important precursor in various biosynthetic pathways. All known FBPases are metal-dependent enzymes, which are classified into five different classes based on their amino acid sequences. Eukaryotes are known to contain only the type-I FBPases, whereas all five types exist in various combinations in prokaryotes. Here we demonstrate that the uncharacterized protein YK23 from Saccharomyces cerevisiae efficiently hydrolyzes FBP in a metal-independent reaction. YK23 is a member of the histidine phosphatase (phosphoglyceromutase) superfamily with homologues found in all organisms. The crystal structure of the YK23 apo-form was solved at 1.75-Å resolution and revealed the core domain with the α/β/α-fold covered by two small cap domains. Two liganded structures of this protein show the presence of two phosphate molecules (an inhibitor) or FBP (a substrate) bound to the active site. FBP is bound in its linear, open conformation with the cleavable C1-phosphate positioned deep in the active site. Alanine replacement mutagenesis of YK23 identified six conserved residues absolutely required for activity and suggested that His13 and Glu99 are the primary catalytic residues. Thus, YK23 represents the first family of metal-independent FBPases and a second FBPase family in eukaryotes.  相似文献   

19.
The intracellular parasite Toxoplasma gondii produces two nucleoside triphosphate diphosphohydrolases (NTPDase1 and -3). These tetrameric, cysteine-rich enzymes require activation by reductive cleavage of a hitherto unknown disulfide bond. Despite a 97% sequence identity, both isozymes differ largely in their ability to hydrolyze ATP and ADP. Here, we present crystal structures of inactive NTPDase3 as an apo form and in complex with the product AMP to resolutions of 2.0 and 2.2 Å, respectively. We find that the enzyme is present in an open conformation that precludes productive substrate binding and catalysis. The cysteine bridge 258–268 is identified to be responsible for locking of activity. Crystal structures of constitutively active variants of NTPDase1 and -3 generated by mutation of Cys258–Cys268 show that opening of the regulatory cysteine bridge induces a pronounced contraction of the whole tetramer. This is accompanied by a 12° domain closure motion resulting in the correct arrangement of all active site residues. A complex structure of activated NTPDase3 with a non-hydrolyzable ATP analog and the cofactor Mg2+ to a resolution of 2.85 Å indicates that catalytic differences between the NTPDases are primarily dictated by differences in positioning of the adenine base caused by substitution of Arg492 and Glu493 in NTPDase1 by glycines in NTPDase3.  相似文献   

20.
Three new structures of Escherichia coli succinate-quinone oxidoreductase (SQR) have been solved. One with the specific quinone-binding site (Q-site) inhibitor carboxin present has been solved at 2.4 Å resolution and reveals how carboxin inhibits the Q-site. The other new structures are with the Q-site inhibitor pentachlorophenol and with an empty Q-site. These structures reveal important details unresolved in earlier structures. Comparison of the new SQR structures shows how subtle rearrangements of the quinone-binding site accommodate the different inhibitors. The position of conserved water molecules near the quinone binding pocket leads to a reassessment of possible water-mediated proton uptake networks that complete reduction of ubiquinone. The dicarboxylate-binding site in the soluble domain of SQR is highly similar to that seen in high resolution structures of avian SQR (PDB 2H88) and soluble flavocytochrome c (PDB 1QJD) showing mechanistically significant structural features conserved across prokaryotic and eukaryotic SQRs.Succinate:quinone oxidoreductase (SQR,4 succinate dehydrogenase) and menaquinol:fumarate oxidoreductase (QFR, fumarate reductase), members of the Complex II family, are homologous integral membrane proteins which couple the interconversion of succinate and fumarate with quinone and quinol (14). SQR is a key enzyme in the Krebs cycle, oxidizing succinate to fumarate during aerobic growth and reducing quinone to quinol and, thus, acts as a direct link between the Krebs cycle and the respiratory chain. QFR is found in anaerobic or facultative bacteria and lower eukaryotes, where it couples the oxidation of reduced quinones to the reduction of fumarate (1, 4). Escherichia coli SQR has four subunits, two hydrophilic subunits exposed to the cytoplasm (SdhA and SdhB), which interact with two hydrophobic membrane-intrinsic subunits (SdhC and SdhD) (5). SdhA contains the dicarboxylate-binding site and a covalently bound FAD cofactor which cycles between FAD and FADH2 redox states during succinate oxidation (6). The electrons from succinate oxidation are sequentially transferred via a [2Fe-2S], a [4Fe-4S], and a [3Fe-4S] iron-sulfur cluster relay system in SdhB to a quinone-binding site (QP) located at the interface of the SdhB, SdhC, and SdhD subunits. SdhC and SdhD are both composed of three transmembrane helices and coordinate a low spin b-type heme via His residues contributed by each subunit (7, 8).The first structural information about members of the Complex II family came from x-ray structures of the QFR enzymes from E. coli at 3.3 Å resolution (9) and Wolinella succinogenes at 2.2 Å resolution (10). These structures revealed details of the overall architecture of the subunits, the position of key redox cofactors, the electron transfer pathway, and the quinone-binding sites. At around the same time, the structures of soluble fumarate reductases found in anaerobic and microaerophilic bacteria and structurally homologous to the flavoprotein subunit of Complex II were solved by x-ray crystallography (1). Analysis of these soluble fumarate reductases has proven particularly informative in describing the mechanism of fumarate reduction and succinate oxidation at the dicarboxylate-binding site (1114).Structures of SQRs lagged behind those of the QFRs until the structure of the E. coli enzyme was solved at 2.6 Å (15). This structure, solved in space group R32, revealed that the E. coli enzyme is packed as a trimer. The structures of the SdhA and SdhB subunits were highly similar to those of E. coli and W. succinogenes QFRs, but the transmembrane SdhC and SdhD subunits showed differences compared with their QFR counterparts. The structure revealed the position of the redox sites and the dicarboxylate- and quinone-binding (Q) sites. The heme b molecule was shown to lie away from the electron transfer pathway, suggesting electrons are preferentially transferred from the [3Fe-4S] cluster to ubiquinone, on the grounds of the edge-to-edge distances and redox potentials of the relevant groups. The structure revealed density in the Q-site that was interpreted as ubiquinone, and the position of the binding site was confirmed by the structure of the E. coli enzyme co-crystallized with the Q-site inhibitor 2-(1-methyl-hexyl)-4,6-dinitrophenol (DNP-17, PDB code 1NEN (15)). The E. coli enzyme was subsequently co-crystallized with the Q-site inhibitor Atpenin A5 (AA5) (PDB code 2ACZ (16)). This inhibitor was bound deeper into the quinone-binding site than ubiquinone or DNP-17, suggesting that there are two binding positions for ubiquinone in its binding site. The structure also identified a water-mediated proton pathway, proposed to deliver protons to the quinone-binding site. The first structure of a mitochondrial SQR was from porcine heart at 2.4 Å resolution (PDB code 1ZOY (17). This structure revealed a monomer in the asymmetric unit, suggesting that mitochondrial SQRs were likely to function as monomers. Superposition of the porcine and E. coli SQR structures revealed the high structural similarity of the SdhA and SdhB subunits and the conservation in position of the redox cofactors. Larger divergences were observed in the transmembrane subunits.Further structural information about SQRs was obtained by analysis of structures of avian SQR crystallized with oxaloacetate (2.2 Å resolution, PDB code 1YQ3), with 3-nitropropionate (2.4 Å resolution, PDB code 1YQ4), and with the Q-site inhibitor carboxin (2.1 Å resolution, PDB code 2FBW) (18). These structures revealed important differences in the position of key residues in the dicarboxylate-binding site compared with the E. coli and porcine structures. Arg-297 (equivalent to Arg-298 in porcine and Arg-286 in E. coli SQRs) was ideally located to act as a general base catalyst, accepting a proton during dehydrogenation of succinate, as in the soluble Shewanella flavocytochrome c3 (PDB code 1QJD) (11), suggesting conservation of mechanism between these distantly related enzymes. An unusual cis-serine peptide bond was proposed to position another arginine residue for binding dicarboxylates. Density for the dicarboxylate in 1YQ3 and 2FBW was shown to be distinctly non-planar and could be modeled by the “malate-like intermediate” seen in 1QJD. The nature of the ligand in the dicarboxylate site was further analyzed in a 1.74 Å resolution structure of avian SQR (PDB code 2H88), confirming the high structural similarity of the ligand and binding site residues in the SQR and flavocytochrome c3 structures (11, 12, 14).Despite the structural information described above, there are still unresolved issues regarding the structure and function of SQRs and QFRs. These include the location of conserved waters, which may form a channel involved in protonation of quinone, and the ability of the Q-site to accommodate different quinones and inhibitors. To further address these issues, we pursued structure-function studies of E. coli SQR. We developed alternative crystallization conditions that provided crystals more reproducibly and diffracting to higher resolution. By exchanging the enzyme into decyl-β-d-maltoside (DM) during purification, it was possible to crystallize the enzyme in the orthorhombic P212121 space group. These crystals routinely diffracted in the 3–3.5 Å resolution range. Co-crystallization with the biochemically well characterized Q-site inhibitor carboxin improved diffraction to 2.1–2.8 Å. This structure shows new features related to the dicarboxylate-binding site of E. coli SQR including a rare cis-peptide bond in SdhA, as found in avian SQR (14), which helps shape the geometry of the active site. Comparisons of the structure with those of SQR binding PCP and SQR with an empty Q-site show how subtle rearrangements of the Q-site accommodate the different inhibitors. The orientation of carboxin in the Q-site differs with computational predictions (16) and with that seen in avian SQR (2FBW). The position of conserved water molecules around the Q-site suggests a new water-mediated proton uptake pathway consistent with recent mutational and biophysical studies (19).  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