1
|
Han X, Qian L, Zhang L, Liu X. Structural and biochemical insights into nucleotide-rhamnose synthase/epimerase-reductase from Arabidopsis thaliana. BIOCHIMICA ET BIOPHYSICA ACTA-PROTEINS AND PROTEOMICS 2015; 1854:1476-86. [PMID: 26116145 DOI: 10.1016/j.bbapap.2015.06.007] [Citation(s) in RCA: 11] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [Key Words] [Subscribe] [Scholar Register] [Received: 02/25/2015] [Revised: 06/03/2015] [Accepted: 06/20/2015] [Indexed: 11/26/2022]
Abstract
L-Rhamnose (Rha) is synthesized via a similar enzymatic pathway in bacteria, plants and fungi. In plants, nucleotide-rhamnose synthase/epimerase-reductase (NRS/ER) catalyzes the final step in the conversion of dTDP/UDP-α-D-Glc to dTDP/UDP-β-L-Rha in an NAD(P)H dependent manner. Currently, only biochemical evidence for the function of NRS/ER has been described. In this study, a crystal structure for Arabidopsis thaliana NRS/ER was determined, which is the first report of a eukaryotic rhamnose synthase with both epimerase and reductase activities. NRS/ER functions as a metal ion independent homodimer that forms through hydrophobic interactions via a four-helix bundle. Each monomer exhibits α/β folding that can be divided into two regions, nucleotide cofactor binding domain and sugar substrate binding domain. The affinities of ligands with NRS/ER were measured using isothermal titration calorimetry, which showed that NRS/ER has a preference for dTDP over UDP, while the cofactor binding site has a similar affinity for NADH and NADPH. Structural analysis coupled to site-directed mutagenesis suggested C115 and K183 as the acid/base pair responsible for epimerization, while T113, Y144 and K148 are the conserved residues in reduction. These findings shed light on the molecular mechanism of NRS/ER and were helpful to explore other eukaryotic enzymes involved in L-Rha synthesis.
Collapse
Affiliation(s)
- Xiaodong Han
- State Key Laboratory of Medicinal Chemical Biology, College of Life Sciences, Nankai University, Tianjin 300071, China; Food and Pharmaceutical Engineering Institute, Shanxi University of Traditional Chinese Medicine, Taiyuan 030024, China.
| | - Lei Qian
- State Key Laboratory of Medicinal Chemical Biology, College of Life Sciences, Nankai University, Tianjin 300071, China; Tianjin Research Institute of Forestry and Pomology, Tianjin Academy of Agricultural Sciences, Tianjin 300192, China.
| | - Lianwen Zhang
- College of Pharmacy, Collaborative Innovation Center for Biotherapy, and Tianjin Key Laboratory of Molecular Drug Research, Nankai University, Tianjin 300071, China.
| | - Xinqi Liu
- State Key Laboratory of Medicinal Chemical Biology, College of Life Sciences, Nankai University, Tianjin 300071, China.
| |
Collapse
|
2
|
Lu DY, Ye J, Han LS, Qiu WJ, Zhang HW, Zhou JD, Bao PZ, Zhang YF, Gu XF. QDPR gene mutation and clinical follow-up in Chinese patients with dihydropteridine reductase deficiency. World J Pediatr 2014; 10:219-26. [PMID: 25124972 DOI: 10.1007/s12519-014-0496-7] [Citation(s) in RCA: 6] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Received: 10/12/2013] [Accepted: 01/10/2014] [Indexed: 11/29/2022]
Abstract
BACKGROUND This study aimed to investigate the mutation spectrum of the QDPR gene, to determine the effect of mutations on dihydropteridine reductase (DHPR) structure/function, to discuss the potential genotypephenotype correlation, and to evaluate the clinical outcome of Chinese patients after treatment. METHODS Nine DHPR-deficient patients were enrolled in this study and seven of them underwent neonatal screening. QDPR gene mutations were analyzed and confirmed by routine methods. The potential pathogenicity of missense variants was analyzed using Clustal X, PolyPhen program and Swiss-PDB Viewer 4.04_OSX software, respectively. The clinical outcomes of the patients were evaluated after long-term treatment. RESULTS In 10 mutations of the 9 patients, 4 were novel mutations (G20V, V86D, G130S and A175R), 4 were reported by us previously, and 2 known mutations were identified. R221X was a hotspot mutation (27.7%) in our patients. Eight missense mutations probably had damage to protein. Six patients in this series were treated with a good control of phenylalanine level. The height and weight of the patients were normal at the age of 4 months to 7.5 years. Four patients, who underwent a neonatal screening and were treated early, showed a normal mental development. In 2 patients diagnosed late, neurological symptoms were significantly improved. CONCLUSIONS The mutation spectrum of the QDPR gene is different in the Chinese population. Most mutations are related to severe phenotype. The determination of DHPR activity should be performed in patients with hyperphenylalaninemia. DHPR-deficient patients who were treated below the age of 2 months may have a near normal mental development.
Collapse
Affiliation(s)
- De-Yun Lu
- Department of Pediatric Endocrinology and Genetic Metabolism and Shanghai Institute for Pediatric Research, Xinhua Hospital, Shanghai Jiaotong University School of Medicine, Shanghai, 200092, China
| | | | | | | | | | | | | | | | | |
Collapse
|
3
|
Chen C, Kim HL, Zhuang N, Seo KH, Park KH, Han CD, Park YS, Lee KH. Structural insights into the dual substrate specificities of mammalian and Dictyostelium dihydropteridine reductases toward two stereoisomers of quinonoid dihydrobiopterin. FEBS Lett 2011; 585:2640-6. [PMID: 21819985 DOI: 10.1016/j.febslet.2011.07.018] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 04/12/2011] [Revised: 07/04/2011] [Accepted: 07/13/2011] [Indexed: 11/25/2022]
Abstract
Up to now, d-threo-tetrahydrobiopterin (DH(4), dictyopterin) was detected only in Dictyostelium discoideum, while the isomer L-erythro-tetrahydrobioterin (BH(4)) is common in mammals. To elucidate the mechanism of DH(4) regeneration by D. discoideum dihydropteridine reductase (DicDHPR), we have determined the crystal structure of DicDHPR complexed with NAD(+) at 2.16 Å resolution. Significant structural differences from mammalian DHPRs are found around the coenzyme binding site, resulting in a higher K(m) value for NADH (K(m)=46.51±0.4 μM) than mammals. In addition, we have found that rat DHPR as well as DicDHPR could bind to both substrates quinonoid-BH(2) and quinonoid-DH(2) by docking calculations and have confirmed their catalytic activity by in vitro assay.
Collapse
Affiliation(s)
- Cong Chen
- Division of Applied Life Science (BK21 Program), Gyeongsang National University, Jinju, Republic of Korea
| | | | | | | | | | | | | | | |
Collapse
|
4
|
Dumitrescu C, Biondi R, Xia Y, Cardounel AJ, Druhan LJ, Ambrosio G, Zweier JL. Myocardial ischemia results in tetrahydrobiopterin (BH4) oxidation with impaired endothelial function ameliorated by BH4. Proc Natl Acad Sci U S A 2007; 104:15081-6. [PMID: 17848522 PMCID: PMC1986616 DOI: 10.1073/pnas.0702986104] [Citation(s) in RCA: 157] [Impact Index Per Article: 9.2] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/30/2007] [Indexed: 02/07/2023] Open
Abstract
Coronary vasodilation is impaired in the postischemic heart with a loss of endothelial nitric oxide synthase (eNOS) activity, but the mechanisms underlying ischemia-induced eNOS dysfunction are not understood. For nitric oxide (NO) synthesis, eNOS requires the redox-sensitive cofactor tetrahydrobiopterin (BH(4)); however, the role of BH(4) in ischemia-induced endothelial dysfunction remains unknown. Therefore, isolated rat hearts were subjected to varying durations of ischemia, and the alterations in NOS-dependent vasodilation were measured and correlated with assays of eNOS activity and cardiac BH(4) concentrations. Ischemia time-dependently decreased cardiac BH(4) content with 85, 95, or 97% irreversible degradation after 30, 45, or 60 min of ischemia, respectively. Paralleling the decreases in BH(4), reductions of eNOS activity were seen of 58, 86, or 92%, and NOS-derived superoxide production was greatly increased. Addition of 10 microM BH(4) enhanced eNOS activity in nonischemic hearts and partially restored activity after ischemia. It also suppressed NOS-derived superoxide production. Impaired coronary flow during postischemic reperfusion was improved by BH(4) infusion. Thus, BH(4) depletion contributes to postischemic eNOS dysfunction, and BH(4) treatment is effective in partial restoration of endothelium-dependent coronary flow. Supplementation of BH(4) may therefore be an important therapeutic approach to reverse endothelial dysfunction in postischemic tissues.
Collapse
Affiliation(s)
- Cristian Dumitrescu
- *Davis Heart and Lung Research Institute, Division of Cardiovascular Medicine, Department of Internal Medicine, College of Medicine, Ohio State University, Columbus, OH 43210; and
| | - Roberto Biondi
- *Davis Heart and Lung Research Institute, Division of Cardiovascular Medicine, Department of Internal Medicine, College of Medicine, Ohio State University, Columbus, OH 43210; and
- Sezione di Immunologia e Patologia Generale, University of Perugia School of Medicine, 06156 Perugia, Italy
| | - Yong Xia
- *Davis Heart and Lung Research Institute, Division of Cardiovascular Medicine, Department of Internal Medicine, College of Medicine, Ohio State University, Columbus, OH 43210; and
| | - Arturo J. Cardounel
- *Davis Heart and Lung Research Institute, Division of Cardiovascular Medicine, Department of Internal Medicine, College of Medicine, Ohio State University, Columbus, OH 43210; and
| | - Lawrence J. Druhan
- *Davis Heart and Lung Research Institute, Division of Cardiovascular Medicine, Department of Internal Medicine, College of Medicine, Ohio State University, Columbus, OH 43210; and
| | - Giuseppe Ambrosio
- *Davis Heart and Lung Research Institute, Division of Cardiovascular Medicine, Department of Internal Medicine, College of Medicine, Ohio State University, Columbus, OH 43210; and
- Division of Cardiology and
| | - Jay L. Zweier
- *Davis Heart and Lung Research Institute, Division of Cardiovascular Medicine, Department of Internal Medicine, College of Medicine, Ohio State University, Columbus, OH 43210; and
| |
Collapse
|
5
|
Gherardini PF, Wass MN, Helmer-Citterich M, Sternberg MJE. Convergent Evolution of Enzyme Active Sites Is not a Rare Phenomenon. J Mol Biol 2007; 372:817-45. [PMID: 17681532 DOI: 10.1016/j.jmb.2007.06.017] [Citation(s) in RCA: 85] [Impact Index Per Article: 5.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 11/16/2006] [Revised: 05/14/2007] [Accepted: 06/08/2007] [Indexed: 02/03/2023]
Abstract
Since convergent evolution of enzyme active sites was first identified in serine proteases, other individual instances of this phenomenon have been documented. However, a systematic analysis assessing the frequency of this phenomenon across enzyme space is still lacking. This work uses the Query3d structural comparison algorithm to integrate for the first time detailed knowledge about catalytic residues, available through the Catalytic Site Atlas (CSA), with the evolutionary information provided by the Structural Classification of Proteins (SCOP) database. This study considers two modes of convergent evolution: (i) mechanistic analogues which are enzymes that use the same mechanism to perform related, but possibly different, reactions (considered here as sharing the first three digits of the EC number); and (ii) transformational analogues which catalyse exactly the same reaction (identical EC numbers), but may use different mechanisms. Mechanistic analogues were identified in 15% (26 out of 169) of the three-digit EC groups considered, showing that this phenomenon is not rare. Furthermore 11 of these groups also contain transformational analogues. The catalytic triad is the most widespread active site; the results of the structural comparison show that this mechanism, or variations thereof, is present in 23 superfamilies. Transformational analogues were identified for 45 of the 951 four-digit EC numbers present within the CSA and about half of these were also mechanistic analogues exhibiting convergence of their active sites. This analysis has also been extended to the whole Protein Data Bank to provide a complete and manually curated list of the all the transformational analogues whose structure is classified in SCOP. The results of this work show that the phenomenon of convergent evolution is not rare, especially when considering large enzymatic families.
Collapse
Affiliation(s)
- Pier Federico Gherardini
- Biochemistry Building, Division of Molecular Biosciences, Imperial College London, London SW7 2AZ, UK
| | | | | | | |
Collapse
|
6
|
Wild C, Golderer G, Gröbner P, Werner-Felmayer G, Werner ER. Physarum polycephalum expresses a dihydropteridine reductase with selectivity for pterin substrates with a 6-(1', 2'-dihydroxypropyl) substitution. Biol Chem 2003; 384:1057-62. [PMID: 12956422 DOI: 10.1515/bc.2003.118] [Citation(s) in RCA: 5] [Impact Index Per Article: 0.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/15/2022]
Abstract
Physarum polycephalum is one of few non-animal organisms capable of synthesizing tetrahydrobiopterin from GTP. Here we demonstrate developmentally regulated expression of quinoid dihydropteridine reductase (EC 1.6.99.7), an enzyme required for recycling 6,7-[8H]-dihydrobiopterin. Physarum also expresses phenylalanine-4-hydroxylase activity, an enzyme that depends on dihydropteridine reductase. The 24.4 kDa Physarum dihydropteridine reductase shares 43% amino acid identity with the human protein. A number of residues important for function of the mammalian enzyme are also conserved in the Physarum sequence. In comparison to sheep liver dihydropteridine reductase, purified recombinant Physarum dihydropteridine reductase prefers pterin substrates with a 6-(1', 2'-dihydroxypropyl) group. Our results demonstrate that Physarum synthesizes, utilizes and metabolizes tetrahydrobiopterin in a way hitherto thought to be restricted to the animal kingdom.
Collapse
Affiliation(s)
- Claudia Wild
- Institute of Medical Chemistry and Biochemistry, University of Innsbruck, Fritz-Pregl-Str. 3, A-6020 Innsbruck, Austria
| | | | | | | | | |
Collapse
|
7
|
Lye LF, Cunningham ML, Beverley SM. Characterization of quinonoid-dihydropteridine reductase (QDPR) from the lower eukaryote Leishmania major. J Biol Chem 2002; 277:38245-53. [PMID: 12151409 DOI: 10.1074/jbc.m206543200] [Citation(s) in RCA: 28] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Biopterin is required for growth of the protozoan parasite Leishmania and is salvaged from the host through the activities of a novel biopterin transporter (BT1) and broad-spectrum pteridine reductase (PTR1). Here we characterize Leishmania major quinonoid-dihydropteridine reductase (LmQDPR), the key enzyme required for regeneration and maintenance of H(4)biopterin pools. LmQDPR shows good homology to metazoan quinonoid-dihydropteridine reductase and conservation of domains implicated in catalysis and regulation. Unlike other organisms, LmQDPR is encoded by a tandemly repeated array of 8-9 copies containing LmQDPR plus two other genes. QDPR mRNA and enzymatic activity were expressed at similar levels throughout the infectious cycle. The pH optima, kinetic properties, and substrate specificity of purified LmQDPR were found to be similar to that of other qDPRs, although it lacked significant activity for non-quinonoid pteridines. These and other data suggest that LmQDPR is unlikely to encode the dihydrobiopterin reductase activity (PTR2) described previously. Similarly LmQDPR is not inhibited by a series of antifolates showing anti-leishmanial activity beyond that attributable to dihydrofolate reductase or PTR1 inhibition. qDPR activity was found in crude lysates of Trypanosoma brucei and Trypanosoma cruzi, further emphasizing the importance of H(4)biopterin throughout this family of human parasites.
Collapse
Affiliation(s)
- Lon-Fye Lye
- Department of Molecular Microbiology, Washington University School of Medicine, St. Louis, Missouri 63110, USA
| | | | | |
Collapse
|
8
|
Park D, Park S, Yim J. Molecular characterization of Drosophila melanogaster dihydropteridine reductase. BIOCHIMICA ET BIOPHYSICA ACTA 2000; 1492:247-51. [PMID: 11004497 DOI: 10.1016/s0167-4781(00)00084-1] [Citation(s) in RCA: 6] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/24/2022]
Abstract
Dihydropteridine reductase (DHPR) catalyzes the NAD(P)H-mediated reduction of quinonoid dihydropteridine as a part of pterin-dependent aromatic amino acid hydroxylation. We isolated a fragment of Drosophila DHPR gene by PCR using degenerate primers. By screening a cDNA library, we obtained full-length clones. The predicted amino acid sequence of the Drosophila DHPR protein was highly homologous to other species including human and mouse. In particular, the Tyr-(Xaa)(3)-Lys motif, known as the NAD(P)H binding domain, and most amino acids relevant to quinonoid dihydropteridine binding site are identical to human DHPR. The recombinant DHPR protein expressed in Escherichia coli showed DHPR enzyme activity. Northern blot analysis revealed two transcripts of 1.1 and 0.9 kb. Genomic DNA sequencing revealed that the gene consists of two exons interrupted by a single 96-bp intron. The two transcripts have alternative promoters, both having no putative TATA box or CAAT box, but sharing a common poly(A)(+) signal. The existence of two alternative promoters suggests that each transcript be regulated independently through different stimuli. Further study is needed to examine the expression and function of the two alternative transcripts.
Collapse
Affiliation(s)
- D Park
- Creative Research Initiative Center, Seoul, South Korea
| | | | | |
Collapse
|
9
|
Somoza JR, Menon S, Schmidt H, Joseph-McCarthy D, Dessen A, Stahl ML, Somers WS, Sullivan FX. Structural and kinetic analysis of Escherichia coli GDP-mannose 4,6 dehydratase provides insights into the enzyme's catalytic mechanism and regulation by GDP-fucose. Structure 2000; 8:123-35. [PMID: 10673432 DOI: 10.1016/s0969-2126(00)00088-5] [Citation(s) in RCA: 72] [Impact Index Per Article: 3.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
BACKGROUND GDP-mannose 4,6 dehydratase (GMD) catalyzes the conversion of GDP-(D)-mannose to GDP-4-keto, 6-deoxy-(D)-mannose. This is the first and regulatory step in the de novo biosynthesis of GDP-(L)-fucose. Fucose forms part of a number of glycoconjugates, including the ABO blood groups and the selectin ligand sialyl Lewis X. Defects in GDP-fucose metabolism have been linked to leukocyte adhesion deficiency type II (LADII). RESULTS The structure of the GDP-mannose 4,6 dehydratase apo enzyme has been determined and refined using data to 2.3 A resolution. GMD is a homodimeric protein with each monomer composed of two domains. The larger N-terminal domain binds the NADP(H) cofactor in a classical Rossmann fold and the C-terminal domain harbors the sugar-nucleotide binding site. We have determined the GMD dissociation constants for NADP, NADPH and GDP-mannose. Each GMD monomer binds one cofactor and one substrate molecule, suggesting that both subunits are catalytically competent. GDP-fucose acts as a competitive inhibitor, suggesting that it binds to the same site as GDP-mannose, providing a mechanism for the feedback inhibition of fucose biosynthesis. CONCLUSIONS The X-ray structure of GMD reveals that it is a member of the short-chain dehydrogenase/reductase (SDR) family of proteins. We have modeled the binding of NADP and GDP-mannose to the enzyme and mutated four of the active-site residues to determine their function. The combined modeling and mutagenesis data suggests that at position 133 threonine substitutes serine as part of the serine-tyrosine-lysine catalytic triad common to the SDR family and Glu 135 functions as an active-site base.
Collapse
Affiliation(s)
- J R Somoza
- Wyeth Research, Cambridge, MA 02140, USA
| | | | | | | | | | | | | | | |
Collapse
|
10
|
Chang CF, Bray T, Whiteley JM. Mutant PTR1 proteins from Leishmania tarentolae: comparative kinetic properties and active-site labeling. Arch Biochem Biophys 1999; 368:161-71. [PMID: 10415124 DOI: 10.1006/abbi.1999.1290] [Citation(s) in RCA: 7] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
PTR1, the gene promoting MTX resistance following gene amplification or DNA transfection in Leishmania tarentolae and selected mutants, has been cloned and heavily overexpressed (>100 mg/liter) in Escherichia coli strain BL21 (DE3). Protein has been purified, essentially to homogeneity, in two steps, via ammonium sulfate precipitation and chromatography on DEAE-Trisacryl. The active proteins are tetramers and display optimal pteridine reductase activity at pH 6.0 using biopterin as substrate and NADPH as the reduced dinucleotide cofactor. 2,4-Diaminopteridine substrate analogues are strong competitive inhibitors (K(i) approximately 38 --> 3 nM) against the pterin substrate and both NADP(+) and folate are inhibitors although somewhat weaker. Dihydropteridines are poor substrates compared to the fully oxidized pteridine. Kinetic analysis affords the usual Michaelis constants and in addition shows that inhibition by NADP(+) allows the formation of ternary nonproductive complexes with folate. The kinetic results are consistent with a sequential ordered bi-bi kinetic mechanism in which first NADPH and then pteridine bind to the free enzyme. Sequence comparisons suggest that PTR1 belongs to the short-chain dehydrogenase/reductase (SDR) family containing an amino-terminal glycine-rich dinucleotide binding site plus a catalytic Y(Xaa)(3)K motif. In accord with this observation, the mutants K16A, Y37D, and R39A and the double mutants K17A:R39A and Y37D:R39A all show a two- to threefold lower binding affinity for NADPH and exhibit low or zero activity. Two Y(Xaa)(3)K regions are present in wild-type PTR1 at 152 and 194. Only Y194F gives protein with zero activity. This observation coupled with affinity labeling of PTR1 by oNADP(+) (2', 3'-dialdehyde derivative of NADP(+)) followed by NaBH(4) reduction, V8 protease digestion, and mass spectral analysis suggests that the motif participating in catalysis is that at 194. The mutation K198Q eliminates inactivation by oNADP(+) supporting the hypothesis that K198 is associated with nucleotide orientation, as has been demonstrated for similar lysine residues in other members of the SDR family.
Collapse
Affiliation(s)
- C F Chang
- The Scripps Research Institute, La Jolla, California, 92037, USA
| | | | | |
Collapse
|
11
|
Somers WS, Stahl ML, Sullivan FX. GDP-fucose synthetase from Escherichia coli: structure of a unique member of the short-chain dehydrogenase/reductase family that catalyzes two distinct reactions at the same active site. Structure 1998; 6:1601-12. [PMID: 9862812 DOI: 10.1016/s0969-2126(98)00157-9] [Citation(s) in RCA: 58] [Impact Index Per Article: 2.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 10/26/2022]
Abstract
BACKGROUND . In all species examined, GDP-fucose is synthesized from GDP-mannose in a three-step reaction catalyzed by two enzymes, GDP-mannose 4,6 dehydratase and a dual function 3, 5-epimerase-4-reductase named GDP-fucose synthetase. In this latter aspect fucose biosynthesis differs from that of other deoxy and dideoxy sugars, in which the epimerase and reductase activities are present as separate enzymes. Defects in GDP-fucose biosynthesis have been shown to affect nodulation in bacteria, stem development in plants, and are associated with the immune defect leukocyte adhesion deficiency type II in humans. RESULTS . We have determined the structure of GDP-fucose synthetase from Escherichia coli at 2.2 A resolution. The structure of GDP-fucose synthetase is closely related to that of UDP-galactose 4-epimerase and more distantly to other members of the short-chain dehydrogenase/reductase family. We have also determined the structures of the binary complexes of GDP-fucose synthetase with its substrate NADPH and its product NADP+. The nicotinamide cofactors bind in the syn and anti conformations, respectively. CONCLUSIONS . GDP-fucose synthetase binds its substrate, NADPH, in the proper orientation (syn) for transferring the 4-pro-S hydride of the nicotinamide. We have observed a single binding site in GDP-fucose synthetase for the second substrate, GDP-4-keto,6-deoxy-mannose. This implies that both the epimerization and reduction reactions occur at the same site in the enzyme. As is the case for all members of the short-chain family of dehydrogenase/reductases, GDP-fucose synthetase retains the Ser-Tyr-Lys catalytic triad. We propose that this catalytic triad functions in a mechanistically equivalent manner in both the epimerization and reduction reactions. Additionally, the X-ray structure has allowed us to identify other residues that are potentially required for substrate binding and catalysis.
Collapse
Affiliation(s)
- W S Somers
- Small Molecule Drug Discovery Genetics Institute, Inc. 87 Cambridgepark Drive, Cambridge, MA 02140, USA
| | | | | |
Collapse
|
12
|
Luba J, Nare B, Liang PH, Anderson KS, Beverley SM, Hardy LW. Leishmania major pteridine reductase 1 belongs to the short chain dehydrogenase family: stereochemical and kinetic evidence. Biochemistry 1998; 37:4093-104. [PMID: 9521731 DOI: 10.1021/bi972693a] [Citation(s) in RCA: 25] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/06/2023]
Abstract
Pteridine reductase 1 (PTR1) is a novel broad spectrum enzyme of pterin and folate metabolism in the protozoan parasite Leishmania. Overexpression of PTR1 confers methotrexate resistance to these protozoa, arising from the enzyme's ability to reduce dihydrofolate and its relative insensitivity to methotrexate. The kinetic mechanism and stereochemical course for the catalyzed reaction confirm PTR1's membership within the short chain dehydrogenase/reductase (SDR) family. With folate as a substrate, PTR1 catalyzes two rounds of reduction, yielding 5,6,7, 8-tetrahydrofolate and oxidizing 2 equiv of NADPH. Dihydrofolate accumulates transiently during folate reduction and is both a substrate and an inhibitor of PTR1. PTR1 transfers the pro-S hydride of NADPH to carbon 6 on the si face of dihydrofolate, producing the same stereoisomer of THF as does dihydrofolate reductase. Product inhibition and isotope partitioning studies support an ordered ternary complex mechanism, with NADPH binding first and NADP+ dissociating after the reduced pteridine. Identical kinetic mechanisms and NAD(P)H hydride chirality preferences are seen with other SDRs. An observed tritium effect upon V/K for reduction of dihydrofolate arising from isotopic substitution of the transferred hydride was suppressed at a high concentration of dihydrofolate, consistent with a steady-state ordered kinetic mechanism. Interestingly, half of the binary enzyme-NADPH complex appears to be incapable of rapid turnover. Fluorescence quenching results also indicate the existence of a nonproductive binary enzyme-dihydrofolate complex. The nonproductive complexes observed between PTR1 and its substrates are unique among members of the SDR family and may provide leads for developing antileishmanial therapeutics.
Collapse
Affiliation(s)
- J Luba
- Department of Pharmacology and Molecular Toxicology, University of Massachusetts Medical Center, Worcester 01605, USA
| | | | | | | | | | | |
Collapse
|
13
|
|
14
|
Kiefer PM, Grimshaw CE, Whiteley JM. The comparative interaction of quinonoid (6R)-dihydrobiopterin and an alternative dihydropterin substrate with wild-type and mutant rat dihydropteridine reductases. Biochemistry 1997; 36:9438-45. [PMID: 9235988 DOI: 10.1021/bi970585i] [Citation(s) in RCA: 7] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023]
Abstract
Kinetic parameters and primary deuterium isotope effects have been determined for wild-type dihydropteridine reductase (EC 1.6.99.7) and the Ala133Ser, Lys150Gln, Tyr146His, Tyr146Phe single, and Tyr146Phe/Ala133Ser and Tyr146Phe/Lys150Gln double mutant enzyme forms using the natural substrate, quinonoid (6R)-l-erythro-dihydrobiopterin (qBH2) and an alternate substrate, quinonoid 6,7-dimethyldihydropteridine (q-6,7-diMePtH2). Mutation at either Tyr146 or Lys150 resulted in pronounced changes in kinetic parameters and isotope effects for both pterin substrates, confirming a critical role for these residues in enzyme-mediated hydride transfer. By contrast, the Ala133Ser mutant was practically indistinguishable from wild-type enzyme. The changes observed, however, were quite different for the two pterin substrates. Thus, kcat for q-6,7-diMePtH2 decreased across the series of mutants from a value of 150 s-1 for wild-type enzyme to essentially zero activity for the Tyr146Phe/Lys150Gln double mutant. Conversely, kcat for qBH2 increased 3-11-fold across the same series of mutants from the wild-type value of 23 s-1. For both pterin substrates, the Km (KPt) increased several orders of magnitude upon mutation of Tyr146 or Lys150, with the greater relative increase using qBH2. Significant primary deuterium isotope effects on kcat (Dkcat) and kcat/KPt (D(kcat/KPt)) observed for the Tyr146 and Lys150 mutants varied depending on the pterin substrate used and ranged up to a maximum value of 5.5-6. For qBH2, where Dkcat < Dkcat/KPt was consistently observed, the rate determining step is ascribed to release of the tetrahydropterin product. For q-6,7-diMePtH2, where in all cases Dkcat = Dkcat/KPt, catalysis is probably limited by an isomerization step occurring prior to hydride transfer. Modeling studies in which qBH2 was docked into the binary E:NADH complex provide a structural rationale for the observed differences between the two pterin substrates. The natural substrate, qBH2, displays a higher affinity for the enzyme active site, presumably due to interaction of the dihydroxypropyl side chain of the substrate with a polar loop of residues containing Asn186, Ser189, and Met190. The location of this loop within the three-dimensional structure is consistent with putative substrate binding loops for other members of the short chain dehydrogenase/reductase (SDR) family, which includes dihydropteridine reductase.
Collapse
Affiliation(s)
- P M Kiefer
- University of California at San Diego, La Jolla, California 92093, USA
| | | | | |
Collapse
|
15
|
Nakanishi M, Kaibe H, Matsuura K, Kakumoto M, Tanaka N, Nonaka T, Mitsui Y, Hara A. Site-directed mutagenesis of residues in coenzyme-binding domain and active site of mouse lung carbonyl reductase. ADVANCES IN EXPERIMENTAL MEDICINE AND BIOLOGY 1997; 414:555-61. [PMID: 9059662 DOI: 10.1007/978-1-4615-5871-2_63] [Citation(s) in RCA: 4] [Impact Index Per Article: 0.1] [Reference Citation Analysis] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 02/03/2023]
Affiliation(s)
- M Nakanishi
- Department of Applied Chemistry, Faculty of Engineering, Gifu University, Japan
| | | | | | | | | | | | | | | |
Collapse
|
16
|
Affiliation(s)
- T. Joseph Kappock
- Department of Chemistry, Yale University, P.O. Box 208107 New Haven, Connecticut 06520-8107
| | | |
Collapse
|