1932

Abstract

2-Oxoglutarate (2OG)-dependent oxygenases (2OGXs) catalyze a remarkably diverse range of oxidative reactions. In animals, these comprise hydroxylations and -demethylations proceeding via hydroxylation; in plants and microbes, they catalyze a wider range including ring formations, rearrangements, desaturations, and halogenations. The catalytic flexibility of 2OGXs is reflected in their biological functions. After pioneering work identified the roles of 2OGXs in collagen biosynthesis, research revealed they also function in plant and animal development, transcriptional regulation, nucleic acid modification/repair, fatty acid metabolism, and secondary metabolite biosynthesis, including of medicinally important antibiotics. In plants, 2OGXs are important agrochemical targets and catalyze herbicide degradation. Human 2OGXs, particularly those regulating transcription, are current therapeutic targets for anemia and cancer. Here, we give an overview of the biochemistry of 2OGXs, providing examples linking to biological function, and outline how knowledge of their enzymology is being exploited in medicine, agrochemistry, and biocatalysis.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-biochem-061516-044724
2018-06-20
2024-05-04
Loading full text...

Full text loading...

/deliver/fulltext/biochem/87/1/annurev-biochem-061516-044724.html?itemId=/content/journals/10.1146/annurev-biochem-061516-044724&mimeType=html&fmt=ahah

Literature Cited

  1. 1.  Myllyharju J. 2015. Collagen hydroxylases. See Ref. 2 149–68
  2. 2.  Hausinger RP, Schofield CJ 2015. 2-Oxoglutarate-Dependent Oxygenases Cambridge, UK: R. Soc. Chem
  3. 3.  Hausinger RP. 2015. Biochemical diversity of 2-oxoglutarate-dependent oxygenases. See Ref. 2 1–58
  4. 4.  Smith JL, Khare D 2015. Recent advances in the structural and mechanistic biology of non-haem Fe(II), 2-oxoglutarate and O2-dependent halogenases. See Ref. 2 401–13
  5. 5.  Hausinger RP. 2004. FeII/α-ketoglutarate-dependent hydroxylases and related enzymes. Crit. Rev. Biochem. Mol. Biol. 39:21–68
    [Google Scholar]
  6. 6.  Walport LJ, Hopkinson RJ, Schofield CJ 2012. Mechanisms of human histone and nucleic acid demethylases. Curr. Opin. Chem. Biol. 16:525–34
    [Google Scholar]
  7. 7.  Aik WS, Chowdhury R, Clifton IJ, Hopkinson RJ, Leissing T et al. 2015. Introduction to structural studies on 2-oxoglutarate-dependent oxygenases and related enzymes. See Ref. 2 59–94
  8. 8.  Clifton IJ, McDonough MA, Ehrismann D, Kershaw NJ, Granatino N, Schofield CJ 2006. Structural studies on 2-oxoglutarate oxygenases and related double-stranded β-helix fold proteins. J. Inorg. Biochem. 100:644–69
    [Google Scholar]
  9. 9.  Markolovic S, Leissing TM, Chowdhury R, Wilkins SE, Lu X, Schofield CJ 2016. Structure-function relationships of human JmjC oxygenases-demethylases versus hydroxylases. Curr. Opin. Struct. Biol. 41:62–72
    [Google Scholar]
  10. 10.  McDonough MA, Loenarz C, Chowdhury R, Clifton IJ, Schofield CJ 2010. Structural studies on human 2-oxoglutarate dependent oxygenases. Curr. Opin. Struct. Biol. 20:659–72
    [Google Scholar]
  11. 11.  Valegard K, van Scheltinga AC, Lloyd MD, Hara T, Ramaswamy S et al. 1998. Structure of a cephalosporin synthase. Nature 394:805–9
    [Google Scholar]
  12. 12.  Elkins JM, Ryle MJ, Clifton IJ, Dunning Hotopp JC, Lloyd JS et al. 2002. X-ray crystal structure of Escherichia coli taurine/α-ketoglutarate dioxygenase complexed to ferrous iron and substrates. Biochemistry 41:5185–92
    [Google Scholar]
  13. 13.  Zhang Z, Ren J, Stammers DK, Baldwin JE, Harlos K, Schofield CJ 2000. Structural origins of the selectivity of the trifunctional oxygenase clavaminic acid synthase. Nat. Struct. Biol. 7:127–33
    [Google Scholar]
  14. 14.  Klose RJ, Kallin EM, Zhang Y 2006. JmjC-domain-containing proteins and histone demethylation. Nat. Rev. Genet. 7:715–27
    [Google Scholar]
  15. 15.  Chowdhury R, Sekirnik R, Brissett NC, Krojer T, Ho CH et al. 2014. Ribosomal oxygenases are structurally conserved from prokaryotes to humans. Nature 510:422–26
    [Google Scholar]
  16. 16.  Aik W, McDonough MA, Thalhammer A, Chowdhury R, Schofield CJ 2012. Role of the jelly-roll fold in substrate binding by 2-oxoglutarate oxygenases. Curr. Opin. Struct. Biol. 22:691–700
    [Google Scholar]
  17. 17.  Horton JR, Upadhyay AK, Qi HH, Zhang X, Shi Y, Cheng X 2010. Enzymatic and structural insights for substrate specificity of a family of jumonji histone lysine demethylases. Nat. Struct. Mol. Biol. 17:38–43
    [Google Scholar]
  18. 18.  Loenarz C, Ge W, Coleman ML, Rose NR, Cooper CD et al. 2010. PHF8, a gene associated with cleft lip/palate and mental retardation, encodes for an Nε-dimethyl lysine demethylase. Hum. Mol. Genet. 19:217–22
    [Google Scholar]
  19. 19.  Anantharajan J, Koski MK, Kursula P, Hieta R, Bergmann U et al. 2013. The structural motifs for substrate binding and dimerization of the α subunit of collagen prolyl 4-hydroxylase. Structure 21:2107–18
    [Google Scholar]
  20. 20.  Hanauske-Abel HM, Gunzler V 1982. A stereochemical concept for the catalytic mechanism of prolylhydroxylase: applicability to classification and design of inhibitors. J. Theor. Biol. 94:421–55
    [Google Scholar]
  21. 21.  Bollinger JM Jr, Chang W-C, Matthews ML, Martinie RJ, Boal AK, Krebs C. 2015. Mechanisms of 2-oxoglutarate-dependent oxygenases: the hydroxylation paradigm and beyond. See Ref. 2 95–122
  22. 22.  Gronke RS, VanDusen WJ, Garsky VM, Jacobs JW, Sardana MK et al. 1989. Aspartyl β-hydroxylase: in vitro hydroxylation of a synthetic peptide based on the structure of the first growth factor-like domain of human factor IX. PNAS 86:3609–13
    [Google Scholar]
  23. 23.  Galonic DP, Vaillancourt FH, Walsh CT 2006. Halogenation of unactivated carbon centers in natural product biosynthesis: trichlorination of leucine during barbamide biosynthesis. J. Am. Chem. Soc. 128:3900–1
    [Google Scholar]
  24. 24.  Pasini D, Cloos PA, Walfridsson J, Olsson L, Bukowski JP et al. 2010. JARID2 regulates binding of the Polycomb repressive complex 2 to target genes in ES cells. Nature 464:306–10
    [Google Scholar]
  25. 25.  Chan MC, Holt-Martyn JP, Schofield CJ, Ratcliffe PJ 2016. Pharmacological targeting of the HIF hydroxylases—a new field in medicine development. Mol. Asp. Med. 47–48:54–75
    [Google Scholar]
  26. 26.  Chowdhury R, Leung IK, Tian YM, Abboud MI, Ge W et al. 2016. Structural basis for oxygen degradation domain selectivity of the HIF prolyl hydroxylases. Nat. Commun. 7:12673
    [Google Scholar]
  27. 27.  Chowdhury R, McDonough MA, Mecinovic J, Loenarz C, Flashman E et al. 2009. Structural basis for binding of hypoxia-inducible factor to the oxygen-sensing prolyl hydroxylases. Structure 17:981–89
    [Google Scholar]
  28. 28.  Ng SS, Kavanagh KL, McDonough MA, Butler D, Pilka ES et al. 2007. Crystal structures of histone demethylase JMJD2A reveal basis for substrate specificity. Nature 448:87–91
    [Google Scholar]
  29. 29.  Price JC, Barr EW, Glass TE, Krebs C, Bollinger JM Jr 2003. Evidence for hydrogen abstraction from C1 of taurine by the high-spin Fe(IV) intermediate detected during oxygen activation by taurine:α-ketoglutarate dioxygenase (TauD). J. Am. Chem. Soc. 125:13008–9
    [Google Scholar]
  30. 30.  Costas M, Mehn MP, Jensen MP, Que L Jr 2004. Dioxygen activation at mononuclear nonheme iron active sites: enzymes, models, and intermediates. Chem. Rev. 104:939–86
    [Google Scholar]
  31. 31.  Mitchell AJ, Dunham NP, Martinie RJ, Bergman JA, Pollock CJ et al. 2017. Visualizing the reaction cycle in an iron(II)- and 2-(oxo)-glutarate-dependent hydroxylase. J. Am. Chem. Soc. 139:13830–36
    [Google Scholar]
  32. 32.  Hamed RB, Gomez-Castellanos JR, Henry L, Ducho C, McDonough MA, Schofield CJ 2013. The enzymes of β-lactam biosynthesis. Nat. Prod. Rep. 30:21–107
    [Google Scholar]
  33. 33.  Tarhonskaya H, Szollossi A, Leung IK, Bush JT, Henry L et al. 2014. Studies on deacetoxycephalosporin C synthase support a consensus mechanism for 2-oxoglutarate dependent oxygenases. Biochemistry 53:2483–93
    [Google Scholar]
  34. 34.  Rose NR, McDonough MA, King ONF, Kawamura A, Schofield CJ 2011. Inhibition of 2-oxoglutarate dependent oxygenases. Chem. Soc. Rev. 40:4364–97
    [Google Scholar]
  35. 35.  Mantri M, Zhang Z, McDonough MA, Schofield CJ 2012. Autocatalysed oxidative modifications to 2-oxoglutarate dependent oxygenases. FEBS J 279:1563–75
    [Google Scholar]
  36. 36.  Rutledge PJ. 2015. Isopenicillin N synthase. See Ref. 2 414–24
  37. 37.  Chowdhury R, Candela-Lena JI, Chan MC, Greenald DJ, Yeoh KK et al. 2013. Selective small molecule probes for the hypoxia inducible factor (HIF) prolyl hydroxylases. ACS Chem. Biol. 8:1488–96
    [Google Scholar]
  38. 38.  Flashman E, Bagg EAL, Chowdhury R, Mecinović J, Loenarz C et al. 2008. Kinetic rationale for selectivity toward N- and C-terminal oxygen-dependent degradation domain substrates mediated by a loop region of hypoxia-inducible factor prolyl hydroxylases. J. Biol. Chem. 283:3808–15
    [Google Scholar]
  39. 39.  Scotti JS, Leung IK, Ge W, Bentley MA, Paps J et al. 2014. Human oxygen sensing may have origins in prokaryotic elongation factor Tu prolyl-hydroxylation. PNAS 111:13331–36
    [Google Scholar]
  40. 40.  Muller TA, Hausinger RP 2015. AlkB and its homologues—DNA repair and beyond. See Ref. 2 246–62
  41. 41.  Bleijlevens B, Shivarattan T, Flashman E, Yang Y, Simpson PJ et al. 2008. Dynamic states of the DNA repair enzyme AlkB regulate product release. EMBO Rep 9:872–77
    [Google Scholar]
  42. 42.  Kelly L, McDonough MA, Coleman ML, Ratcliffe PJ, Schofield CJ 2009. Asparagine β-hydroxylation stabilizes the ankyrin repeat domain fold. Mol. Biosyst. 5:52–58
    [Google Scholar]
  43. 43.  Schofield CJ, Ratcliffe PJ 2005. Signalling hypoxia by HIF hydroxylases. Biochem. Biophys. Res. Commun. 338:617–26
    [Google Scholar]
  44. 44.  Chin RM, Fu X, Pai MY, Vergnes L, Hwang H et al. 2014. The metabolite α-ketoglutarate extends lifespan by inhibiting ATP synthase and TOR. Nature 510:397–401
    [Google Scholar]
  45. 45.  Nowicki S, Gottlieb E 2015. Oncometabolites: tailoring our genes. FEBS J 282:2796–805
    [Google Scholar]
  46. 46.  Lindstedt G, Lindstedt S, Nordin I 1982. γ-Butyrobetaine hydroxylase in human kidney. Scand. J. Clin. Lab. Invest. 42:477–85
    [Google Scholar]
  47. 47.  Vaz FM, van Vlies N 2015. Dioxygenases of carnitine biosynthesis: 6-N-trimethyllysine and γ-butyrobetaine hydroxylases. See Ref. 2 324–37
  48. 48.  Berk L, Burchenal JH, Castle WB 1949. Erythropoietic effect of cobalt in patients with or without anemia. N. Engl. J. Med. 240:754–61
    [Google Scholar]
  49. 49.  Salahudeen AA, Thompson JW, Ruiz JC, Ma H-W, Kinch LN et al. 2009. An E3 ligase possessing an iron-responsive hemerythrin domain is a regulator of iron homeostasis. Science 326:722–26
    [Google Scholar]
  50. 50.  Clarke HT. 1952. Henry Drysdale Dakin, 1880–1952. J. Biol. Chem. 198:491–94
    [Google Scholar]
  51. 51.  Ruotsalainen H, Risteli M, Wang C, Wang Y, Karppinen M et al. 2012. The activities of lysyl hydroxylase 3 (LH3) regulate the amount and oligomerization status of adiponectin. PLOS ONE 7:e50045
    [Google Scholar]
  52. 52.  Shoulders MD, Raines RT 2009. Collagen structure and stability. Annu. Rev. Biochem. 78:929–58
    [Google Scholar]
  53. 53.  Aihara A, Huang CK, Olsen MJ, Lin Q, Chung W et al. 2014. A cell-surface β-hydroxylase is a biomarker and therapeutic target for hepatocellular carcinoma. Hepatology 60:1302–13
    [Google Scholar]
  54. 54.  Patel N, Khan AO, Mansour A, Mohamed JY, Al-Assiri A et al. 2014. Mutations in ASPH cause facial dysmorphism, lens dislocation, anterior-segment abnormalities, and spontaneous filtering blebs, or Traboulsi syndrome. Am. J. Hum. Gen. 94:755–59
    [Google Scholar]
  55. 55.  Loenarz C, Schofield CJ 2011. Physiological and biochemical aspects of hydroxylations and demethylations catalyzed by human 2-oxoglutarate oxygenases. Trends. Biochem. Sci. 36:7–18
    [Google Scholar]
  56. 56.  Jiang BH, Rue E, Wang GL, Roe R, Semenza GL 1996. Dimerization, DNA binding, and transactivation properties of hypoxia-inducible factor 1. J. Biol. Chem. 271:17771–78
    [Google Scholar]
  57. 57.  Chowdhury R, Hardy A, Schofield CJ 2008. The human oxygen sensing machinery and its manipulation. Chem. Soc. Rev. 37:1308–19
    [Google Scholar]
  58. 58.  Kaelin WG Jr, Ratcliffe PJ. 2008. Oxygen sensing by metazoans: the central role of the HIF hydroxylase pathway. Mol. Cell 30:393–402
    [Google Scholar]
  59. 59.  Semenza GL, Wang GL 1992. A nuclear factor induced by hypoxia via de novo protein synthesis binds to the human erythropoietin gene enhancer at a site required for transcriptional activation. Mol. Cell Biol. 12:5447–54
    [Google Scholar]
  60. 60.  Wang GL, Jiang BH, Rue EA, Semenza GL 1995. Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS heterodimer regulated by cellular O2 tension. PNAS 92:5510–14
    [Google Scholar]
  61. 61.  Bruick RK, McKnight SL 2001. A conserved family of prolyl-4-hydroxylases that modify HIF. Science 294:1337–40
    [Google Scholar]
  62. 62.  Epstein AC, Gleadle JM, McNeill LA, Hewitson KS, O'Rourke J et al. 2001. C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell 107:43–54
    [Google Scholar]
  63. 63.  Ivan M, Kondo K, Yang H, Kim W, Valiando J et al. 2001. HIFα targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science 292:464–68
    [Google Scholar]
  64. 64.  Jaakkola P, Mole DR, Tian YM, Wilson MI, Gielbert J et al. 2001. Targeting of HIF-α to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 292:468–72
    [Google Scholar]
  65. 65.  Hewitson KS, McNeill LA, Riordan MV, Tian YM, Bullock AN et al. 2002. Hypoxia-inducible factor (HIF) asparagine hydroxylase is identical to factor inhibiting HIF (FIH) and is related to the cupin structural family. J. Biol. Chem. 277:26351–55
    [Google Scholar]
  66. 66.  Lando D, Peet DJ, Gorman JJ, Whelan DA, Whitelaw ML, Bruick RK 2002. FIH-1 is an asparaginyl hydroxylase enzyme that regulates the transcriptional activity of hypoxia-inducible factor. Genes Dev 16:1466–71
    [Google Scholar]
  67. 67.  Chowdhury R, Leung IKH, Tian Y-M, Abboud MI, Ge W et al. 2016. Structural basis for oxygen degradation domain selectivity of the HIF prolyl hydroxylases. Nat. Commun. 7:12673
    [Google Scholar]
  68. 68.  Loenarz C, Coleman ML, Boleininger A, Schierwater B, Holland PW et al. 2011. The hypoxia-inducible transcription factor pathway regulates oxygen sensing in the simplest animal. Trichoplax adhaerens. EMBO Rep. 12:63–70
    [Google Scholar]
  69. 69.  West CM, Blader IJ 2015. Oxygen sensing by protozoans: how they catch their breath. Curr. Opin. Microbiol. 26:41–47
    [Google Scholar]
  70. 70.  Chan MC, Ilott NE, Schödel J, Sims D, Tumber A et al. 2016. Tuning the transcriptional response to hypoxia by inhibiting HIF prolyl- and asparaginyl-hydroxylases. J. Biol. Chem. 291:20661–73
    [Google Scholar]
  71. 71.  Flashman E, Hoffart LM, Hamed RB, Bollinger JM Jr, Krebs C, Schofield CJ 2010. Evidence for the slow reaction of hypoxia-inducible factor prolyl hydroxylase 2 with oxygen. FEBS J 277:4089–99
    [Google Scholar]
  72. 72.  Tarhonskaya H, Hardy AP, Howe EA, Loik ND, Kramer HB et al. 2015. Kinetic investigations of the role of factor inhibiting hypoxia-inducible factor (FIH) as an oxygen sensor. J. Biol. Chem. 290:19726–42
    [Google Scholar]
  73. 73.  Tian Y-M, Yeoh KK, Lee MK, Eriksson T, Kessler BM et al. 2011. Differential sensitivity of hypoxia inducible factor hydroxylation sites to hypoxia and hydroxylase inhibitors. J. Biol. Chem. 286:13041–51
    [Google Scholar]
  74. 74.  Cockman ME, Webb JD, Kramer HB, Kessler BM, Ratcliffe PJ 2009. Proteomics-based identification of novel factor inhibiting hypoxia-inducible factor (FIH) substrates indicates widespread asparaginyl hydroxylation of ankyrin repeat domain-containing proteins. Mol. Cell Proteom. 8:535–46
    [Google Scholar]
  75. 75.  Hardy AP, Prokes I, Kelly L, Campbell ID, Schofield CJ 2009. Asparaginyl β-hydroxylation of proteins containing ankyrin repeat domains influences their stability and function. J. Mol. Biol. 392:994–1006
    [Google Scholar]
  76. 76.  Schmierer B, Novak B, Schofield CJ 2010. Hypoxia-dependent sequestration of an oxygen sensor by a widespread structural motif can shape the hypoxic response—a predictive kinetic model. BMC Syst. Biol. 4:139
    [Google Scholar]
  77. 77.  Loenarz C, Sekirnik R, Thalhammer A, Ge W, Spivakovsky E et al. 2014. Hydroxylation of the eukaryotic ribosomal decoding center affects translational accuracy. PNAS 111:4019–24
    [Google Scholar]
  78. 78.  Singleton RS, Liu-Yi P, Formenti F, Ge W, Sekirnik R et al. 2014. OGFOD1 catalyzes prolyl hydroxylation of RPS23 and is involved in translation control and stress granule formation. PNAS 111:4031–36
    [Google Scholar]
  79. 79.  Horita S, Scotti JS, Thinnes C, Mottaghi-Taromsari YS, Thalhammer A et al. 2015. Structure of the ribosomal oxygenase OGFOD1 provides insights into the regio- and stereoselectivity of prolyl hydroxylases. Structure 23:639–52
    [Google Scholar]
  80. 80.  Katz MJ, Acevedo JM, Loenarz C, Galagovsky D, Liu-Yi P et al. 2014. Sudestada1, a Drosophila ribosomal prolyl-hydroxylase required for mRNA translation, cell homeostasis, and organ growth. PNAS 111:4025–30
    [Google Scholar]
  81. 81.  Bottger A, Islam MS, Chowdhury R, Schofield CJ, Wolf A 2015. The oxygenase Jmjd6—a case study in conflicting assignments. Biochem. J. 468:191–202
    [Google Scholar]
  82. 82.  Chang B, Chen Y, Zhao Y, Bruick RK 2007. JMJD6 is a histone arginine demethylase. Science 318:444–47
    [Google Scholar]
  83. 83.  Ge W, Wolf A, Feng T, Ho CH, Sekirnik R et al. 2012. Oxygenase-catalyzed ribosome hydroxylation occurs in prokaryotes and humans. Nat. Chem. Biol. 8:960–62
    [Google Scholar]
  84. 84.  van Staalduinen LM, Novakowski SK, Jia Z 2014. Structure and functional analysis of YcfD, a novel 2-oxoglutarate/Fe2+-dependent oxygenase involved in translational regulation in Escherichia coli. J. Mol. Biol. 426:1898–910
    [Google Scholar]
  85. 85.  Feng T, Yamamoto A, Wilkins SE, Sokolova E, Yates LA et al. 2014. Optimal translational termination requires C4 lysyl hydroxylation of eRF1. Mol. Cell 53:645–54
    [Google Scholar]
  86. 86.  Del Rizzo PA, Krishnan S, Trievel RC 2012. Crystal structure and functional analysis of JMJD5 indicate an alternate specificity and function. Mol. Cell Biol. 32:4044–52
    [Google Scholar]
  87. 87.  Cheng X, Trievel RC 2015. JmjC lysine demethylases. See Ref. 2 210–45
  88. 88.  Couture JF, Trievel RC 2006. Histone-modifying enzymes: encrypting an enigmatic epigenetic code. Curr. Opin. Struct. Biol. 16:753–60
    [Google Scholar]
  89. 89.  Thinnes CC, England KS, Kawamura A, Chowdhury R, Schofield CJ, Hopkinson RJ 2014. Targeting histone lysine demethylases—progress, challenges, and the future. Biochim. Biophys. Acta 1839:1416–32
    [Google Scholar]
  90. 90.  Walport LJ, Hopkinson RJ, Chowdhury R, Schiller R, Ge W et al. 2016. Arginine demethylation is catalysed by a subset of JmjC histone lysine demethylases. Nat. Commun. 7:11974
    [Google Scholar]
  91. 91.  Horton JR, Upadhyay AK, Hashimoto H, Zhang X, Cheng X 2011. Structural basis for human PHF2 Jumonji domain interaction with metal ions. J. Mol. Biol. 406:1–8
    [Google Scholar]
  92. 92.  Peng JC, Valouev A, Swigut T, Zhang J, Zhao Y et al. 2009. Jarid2/Jumonji coordinates control of PRC2 enzymatic activity and target gene occupancy in pluripotent cells. Cell 139:1290–302
    [Google Scholar]
  93. 93.  Tsukada Y, Ishitani T, Nakayama KI 2010. KDM7 is a dual demethylase for histone H3 Lys 9 and Lys 27 and functions in brain development. Genes Dev 24:432–37
    [Google Scholar]
  94. 94.  Zheng G, He C 2015. RNA demethylation by FTO and ALKBH5. See Ref. 2 263–74
  95. 95.  Falnes PO, Yen Ho AY 2015. Role of ALKBH8 in the synthesis of wobble uridine modifications in tRNA. See Ref. 2 275–88
  96. 96.  Aravind L, Zhang D, Iyer LM 2015. The TET/JBP family of nucleic acid base-modifying 2-oxoglutarate and iron-dependent dioxygenases. See Ref. 2 289–308
  97. 97.  Reynolds D, Cliffe L, Sabatini R 2015. 2-Oxoglutarate-dependent hydroxylases involved in DNA Base J (β-d-glucopyranosyloxymethyluracil) synthesis. See Ref. 2 309–23
  98. 98.  Thornburg LD, Stubbe J 1993. Mechanism-based inactivation of thymine hydroxylase, an α-ketoglutarate-dependent dioxygenase, by 5-ethynyluracil. Biochemistry 32:14034–42
    [Google Scholar]
  99. 99.  Chen K, Zhao BS, He C 2016. Nucleic acid modifications in regulation of gene expression. Cell Chem. Biol. 23:74–85
    [Google Scholar]
  100. 100.  Ringvoll J, Nordstrand LM, Vågbø CB, Talstad V, Reite K et al. 2006. Repair deficient mice reveal mABH2 as the primary oxidative demethylase for repairing 1meA and 3meC lesions in DNA. EMBO J 25:2189–98
    [Google Scholar]
  101. 101.  Duncan T, Trewick SC, Koivisto P, Bates PA, Lindahl T, Sedgwick B 2002. Reversal of DNA alkylation damage by two human dioxygenases. PNAS 99:16660–65
    [Google Scholar]
  102. 102.  Kato M, Araiso Y, Noma A, Nagao A, Suzuki T et al. 2011. Crystal structure of a novel JmjC-domain-containing protein, TYW5, involved in tRNA modification. Nucleic Acids Res 39:1576–85
    [Google Scholar]
  103. 103.  Noma A, Ishitani R, Kato M, Nagao A, Nureki O, Suzuki T 2010. Expanding role of the Jumonji C domain as an RNA hydroxylase. J. Biol. Chem. 285:34503–7
    [Google Scholar]
  104. 104.  Fu Y, Dai Q, Zhang W, Ren J, Pan T, He C 2010. The AlkB domain of mammalian ABH8 catalyzes hydroxylation of 5-methoxycarbonylmethyluridine at the wobble position of tRNA. Angew. Chem. Int. Ed. 49:8885–88
    [Google Scholar]
  105. 105.  van den Born E, Vagbo CB, Songe-Moller L, Leihne V, Lien GF et al. 2011. ALKBH8-mediated formation of a novel diastereomeric pair of wobble nucleosides in mammalian tRNA. Nat. Commun. 2:172
    [Google Scholar]
  106. 106.  He YF, Li BZ, Li Z, Liu P, Wang Y et al. 2011. Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science 333:1303–7
    [Google Scholar]
  107. 107.  Ito S, Shen L, Dai Q, Wu SC, Collins LB et al. 2011. Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science 333:1300–3
    [Google Scholar]
  108. 108.  Pfaffeneder T, Hackner B, Truss M, Munzel M, Muller M et al. 2011. The discovery of 5-formylcytosine in embryonic stem cell DNA. Angew. Chem. Int. Ed. 50:7008–12
    [Google Scholar]
  109. 109.  Tahiliani M, Koh KP, Shen Y, Pastor WA, Bandukwala H et al. 2009. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 324:930–35
    [Google Scholar]
  110. 110.  Cliffe LJ, Hirsch G, Wang J, Ekanayake D, Bullard W et al. 2012. JBP1 and JBP2 proteins are Fe2+/2-oxoglutarate-dependent dioxygenases regulating hydroxylation of thymidine residues in trypanosome DNA. J. Biol. Chem. 287:19886–95
    [Google Scholar]
  111. 111.  Hu L, Li Z, Cheng J, Rao Q, Gong W et al. 2013. Crystal structure of TET2-DNA complex: insight into TET-mediated 5mC oxidation. Cell 155:1545–55
    [Google Scholar]
  112. 112.  Hara R, Kino K 2009. Characterization of novel 2-oxoglutarate dependent dioxygenases converting l-proline to cis-4-hydroxy-l-proline. Biochem. Biophys. Res. Commun. 379:882–86
    [Google Scholar]
  113. 113.  Lawrence CC, Sobey WJ, Field RA, Baldwin JE, Schofield CJ 1996. Purification and initial characterization of proline 4-hydroxylase from Streptomyces griseoviridus P8648: a 2-oxoacid, ferrous-dependent dioxygenase involved in etamycin biosynthesis. Biochem. J. 313:185–91
    [Google Scholar]
  114. 114.  Strieker M, Kopp F, Mahlert C, Essen LO, Marahiel MA 2007. Mechanistic and structural basis of stereospecific Cβ-hydroxylation in calcium-dependent antibiotic, a daptomycin-type lipopeptide. ACS Chem. Biol. 2:187–96
    [Google Scholar]
  115. 115.  Singh GM, Fortin PD, Koglin A, Walsh CT 2008. Beta hydroxylation of the aspartyl residue in the phytotoxin syringomycin E: characterization of two candidate hydroxylases AspH and SyrP in Pseudomonas syringae. Biochemistry 47:11310–20
    [Google Scholar]
  116. 116.  Yin X, Zabriskie TM 2004. VioC is a non-heme iron, α-ketoglutarate-dependent oxygenase that catalyzes the formation of 3S-hydroxy-l-arginine during viomycin biosynthesis. ChemBioChem 5:1274–77
    [Google Scholar]
  117. 117.  Smirnov SV, Sokolov PM, Kodera T, Sugiyama M, Hibi M et al. 2012. A novel family of bacterial dioxygenases that catalyse the hydroxylation of free amino acids. FEMS Microbiol. Lett. 331:97–104
    [Google Scholar]
  118. 118.  Smirnov SV, Kodera T, Samsonova NN, Kotlyarova VA, Rushkevich NY et al. 2010. Metabolic engineering of Escherichia coli to produce (2S, 3R, 4S)-4-hydroxyisoleucine. Appl. Microbiol. Biotechnol. 88:719–26
    [Google Scholar]
  119. 119.  Smirnov SV, Sokolov PM, Kotlyarova VA, Samsonova NN, Kodera T et al. 2013. A novel L-isoleucine-4′-dioxygenase and l-isoleucine dihydroxylation cascade in Pantoea ananatis. MicrobiologyOpen 2:471–81
    [Google Scholar]
  120. 120.  Haefele C, Bonfils C, Sauvaire Y 1997. Characterization of a dioxygenase from Trigonella foenum-graecum involved in 4-hydroxyisoleucine biosynthesis. Phytochemistry 44:563–66
    [Google Scholar]
  121. 121.  Novak RF, Swift TJ, Hoppel CL 1980. N6-Trimethyl-lysine metabolism. Structural identification of the metabolite 3-hydroxy-N6-trimethyl-lysine. Biochem. J. 188:521–27
    [Google Scholar]
  122. 122.  Haltli B, Tan Y, Magarvey NA, Wagenaar M, Yin X et al. 2005. Investigating β-hydroxyenduracididine formation in the biosynthesis of the mannopeptimycins. Chem. Biol. 12:1163–68
    [Google Scholar]
  123. 123.  Bursy J, Pierik AJ, Pica N, Bremer E 2007. Osmotically induced synthesis of the compatible solute hydroxyectoine is mediated by an evolutionarily conserved ectoine hydroxylase. J. Biol. Chem. 282:31147–55
    [Google Scholar]
  124. 124.  Fukuda H, Ogawa T, Tazaki M, Nagahama K, Fujii T et al. 1992. Two reactions are simultaneously catalyzed by a single enzyme: the arginine-dependent simultaneous formation of two products, ethylene and succinate, from 2-oxoglutarate by an enzyme from Pseudomonas syringae. Biochem. Biophys. Res. Commun. 188:483–89
    [Google Scholar]
  125. 125.  Martinez S, Hausinger RP 2016. Biochemical and spectroscopic characterization of the non-heme Fe(II)- and 2-oxoglutarate-dependent ethylene-forming enzyme from Pseudomonas syringae pv. phaseolicola PK2. Biochemistry 55:5989–99
    [Google Scholar]
  126. 126.  Townsend CA. 1997. Structural studies of natural product biosynthetic proteins. Chem. Biol. 4:721–30
    [Google Scholar]
  127. 127.  Chang W-C, Guo Y, Wang C, Butch SE, Rosenzweig AC et al. 2014. Mechanism of the C5 stereoinversion reaction in the biosynthesis of carbapenem antibiotics. Science 343:1140–44
    [Google Scholar]
  128. 128.  Blasiak LC, Vaillancourt FH, Walsh CT, Drennan CL 2006. Crystal structure of the non-haem iron halogenase SyrB2 in syringomycin biosynthesis. Nature 440:368–71
    [Google Scholar]
  129. 129.  Mitchell AJ, Zhu Q, Maggiolo AO, Ananth NR, Hillwig ML et al. 2016. Structural basis for halogenation by iron- and 2-oxo-glutarate-dependent enzyme WelO5. Nat. Chem. Biol. 12:636–40
    [Google Scholar]
  130. 130.  Hedden P, Phillips AL 2015. Gibberellin metabolism. See Ref. 2 367–84
  131. 131.  Martens S, Matern U 2015. Role of 2-oxoglutarate-dependent oxygenases in flavonoid metabolism. See Ref. 2 350–66
  132. 132.  Peiser GD, Wang TT, Hoffman NE, Yang SF, Liu HW, Walsh CT 1984. Formation of cyanide from carbon 1 of 1-aminocyclopropane-1-carboxylic acid during its conversion to ethylene. PNAS 81:3059–63
    [Google Scholar]
  133. 133.  Zhang Z, Ren JS, Clifton IJ, Schofield CJ 2004. Crystal structure and mechanistic implications of 1-aminocyclopropane-1-carboxylic acid oxidase—the ethylene-forming enzyme. Chem. Biol. 11:1383–94
    [Google Scholar]
  134. 134.  Fukumori F, Hausinger RP 1993. Alcaligenes eutrophus JMP134 “2,4-dichlorophenoxyacetate monooxygenase” is an alpha-ketoglutarate-dependent dioxygenase. J. Bacteriol. 175:2083–86
    [Google Scholar]
  135. 135.  Jakubczyk D, Caputi L, Hatsch A, Nielsen CA, Diefenbacher M et al. 2015. Discovery and reconstitution of the cycloclavine biosynthetic pathway—enzymatic formation of a cyclopropyl group. Angew. Chem. Int. Ed. 54:5117–21
    [Google Scholar]
  136. 136.  Wanders RJA, Ferdinandusse S, Ebberink MS, Waterham HR 2015. Phytanoyl-CoA hydroxylase: a 2-oxoglutarate-dependent dioxygenase crucial for fatty acid alpha-oxidation in humans. See Ref. 2 338–49
  137. 137.  Tars K, Leitans J, Kazaks A, Zelencova D, Liepinsh E et al. 2014. Targeting carnitine biosynthesis: discovery of new inhibitors against γ-butyrobetaine hydroxylase. J. Med. Chem. 57:2213–36
    [Google Scholar]
  138. 138.  Vinogradova M, Gehling VS, Gustafson A, Arora S, Tindell CA et al. 2016. An inhibitor of KDM5 demethylases reduces survival of drug-tolerant cancer cells. Nat. Chem. Biol. 12:531–38
    [Google Scholar]
  139. 139.  Henry L, Leung IK, Claridge TD, Schofield CJ 2012. γ-Butyrobetaine hydroxylase catalyses a Stevens type rearrangement. Bioorg. Med. Chem. Lett. 22:4975–78
    [Google Scholar]
  140. 140.  Leung IK, Krojer TJ, Kochan GT, Henry L, von Delft F et al. 2010. Structural and mechanistic studies on γ-butyrobetaine hydroxylase. Chem. Biol. 17:1316–24
    [Google Scholar]
  141. 141.  Rydzik AM, Chowdhury R, Kochan GT, Williams ST, McDonough MA et al. 2014. Modulating carnitine levels by targeting its biosynthesis pathway—selective inhibition of γ-butyrobetaine hydroxylase. Chem. Sci. 5:1765–71
    [Google Scholar]
  142. 142.  Leung IK, Demetriades M, Hardy AP, Lejeune C, Smart TJ et al. 2013. Reporter ligand NMR screening method for 2-oxoglutarate oxygenase inhibitors. J. Med. Chem. 56:547–55
    [Google Scholar]
  143. 143.  Rydzik AM, Leung IK, Kochan GT, Thalhammer A, Oppermann U et al. 2012. Development and application of a fluoride-detection-based fluorescence assay for γ-butyrobetaine hydroxylase. ChemBioChem 13:1559–63
    [Google Scholar]
  144. 144.  Hopkinson RJ, Tumber A, Yapp C, Chowdhury R, Aik W et al. 2013. 5-Carboxy-8-hydroxyquinoline is a broad spectrum 2-oxoglutarate oxygenase inhibitor which causes iron translocation. Chem. Sci. 4:3110–17
    [Google Scholar]
  145. 145.  Rabinowitz MH. 2013. Inhibition of hypoxia-inducible factor prolyl hydroxylase domain oxygen sensors: tricking the body into mounting orchestrated survival and repair responses. J. Med. Chem. 56:9369–402
    [Google Scholar]
  146. 146.  Wilkins SE, Flashman E, Scotti JS, Hopkinson RJ, Chowdhury R, Schofield CJ 2015. The role of 2-oxoglutarate-dependent oxygenases in hypoxia sensing. See Ref. 2 169–209
  147. 147.  Chan MC, Atasoylu O, Hodson E, Tumber A, Leung IK et al. 2015. Potent and selective triazole-based inhibitors of the hypoxia-inducible factor prolyl-hydroxylases with activity in the murine brain. PLOS ONE 10:e0132004
    [Google Scholar]
  148. 148.  Schuster SJ, Badiavas EV, Costa-Giomi P, Weinmann R, Erslev AJ, Caro J 1989. Stimulation of erythropoietin gene transcription during hypoxia and cobalt exposure. Blood 73:13–16
    [Google Scholar]
  149. 149.  Wolf J, Levy IJ 1954. Treatment of sickle cell anemia with cobalt chloride. AMA Arch. Intern. Med. 93:387–96
    [Google Scholar]
  150. 150.  Rademacher W. 2000. Growth retardants: effects on gibberellin biosynthesis and other metabolic pathways. Annu. Rev. Plant Physiol. Plant Mol. Biol. 51:501–31
    [Google Scholar]
  151. 151.  Hedden P. 2003. The genes of the Green Revolution. Trends Genet 19:5–9
    [Google Scholar]
  152. 152.  Rose NR, Woon EC, Tumber A, Walport LJ, Chowdhury R et al. 2012. Plant growth regulator daminozide is a selective inhibitor of human KDM2/7 histone demethylases. J. Med. Chem. 55:6639–43
    [Google Scholar]
  153. 153.  Shah DD, Moran GR 2015. 4-Hydroxyphenylpyruvate dioxygenase and hydroxymandelate synthase: 2-oxo acid-dependent oxygenases of importance to agriculture and medicine. See Ref. 2 438–57
  154. 154.  Lindstedt S, Holme E, Lock EA, Hjalmarson O, Strandvik B 1992. Treatment of hereditary tyrosinaemia type I by inhibition of 4-hydroxyphenylpyruvate dioxygenase. Lancet 340:813–17
    [Google Scholar]
  155. 155.  McAllister TE, England KS, Hopkinson RJ, Brennan PE, Kawamura A, Schofield CJ 2016. Recent progress in histone demethylase inhibitors. J. Med. Chem. 59:1308–29
    [Google Scholar]
  156. 156.  Kawamura A, Münzel M, Kojima T, Yapp C, Bhushan B et al. 2017. Highly selective inhibition of histone demethylases by de novo macrocyclic peptides. Nature Commun 8:14773
    [Google Scholar]
  157. 157.  Chowdhury R, Yeoh KK, Tian YM, Hillringhaus L, Bagg EA et al. 2011. The oncometabolite 2-hydroxyglutarate inhibits histone lysine demethylases. EMBO Rep 12:463–69
    [Google Scholar]
  158. 158.  Flavahan WA, Drier Y, Liau BB, Gillespie SM, Venteicher AS et al. 2016. Insulator dysfunction and oncogene activation in IDH mutant gliomas. Nature 529:110–14
    [Google Scholar]
  159. 159.  Xu W, Yang H, Liu Y, Yang Y, Wang P et al. 2011. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of α-ketoglutarate-dependent dioxygenases. Cancer Cell 19:17–30
    [Google Scholar]
  160. 160.  Koivunen P, Lee S, Duncan CG, Lopez G, Lu G et al. 2012. Transformation by the (R)-enantiomer of 2-hydroxyglutarate linked to EGLN activation. Nature 483:484–88
    [Google Scholar]
  161. 161.  Tarhonskaya H, Rydzik AM, Leung IK, Loik ND, Chan MC et al. 2014. Non-enzymatic chemistry enables 2-hydroxyglutarate-mediated activation of 2-oxoglutarate oxygenases. Nat. Commun. 5:3423
    [Google Scholar]
  162. 162.  Shibasaki T, Mori H, Ozaki A 2000. Enzymatic production of trans-4-hydroxy-l-proline by regio- and stereospecific hydroxylation of l-proline. Biosci. Biotechnol. Biochem. 64:746–50
    [Google Scholar]
  163. 163.  Zafar MI, Gao F 2016. 4-Hydroxyisoleucine: a potential new treatment for type 2 diabetes mellitus. BioDrugs 30:255–62
    [Google Scholar]
  164. 164.  Lin B, Fan K, Zhao J, Ji J, Wu L et al. 2015. Reconstitution of TCA cycle with DAOCS to engineer Escherichia coli into an efficient whole cell catalyst of penicillin G. PNAS 112:9855–59
    [Google Scholar]
  165. 165.  Queener SW. 1990. Molecular biology of penicillin and cephalosporin biosynthesis. Antimicrob. Agents Chemother. 34:943–48
    [Google Scholar]
  166. 166.  Balakrishnan N, Ganesan S, Rajasekaran P, Rajendran L, Teddu S, Durairaaj M 2016. Modified deacetylcephalosporin C synthase for the biotransformation of semisynthetic cephalosporins. Appl. Environ. Microbiol. 82:3711–20
    [Google Scholar]
  167. 167.  Yang M, Hardy AP, Chowdhury R, Loik ND, Scotti JS et al. 2013. Substrate selectivity analyses of factor inhibiting hypoxia-inducible factor. Angew. Chem. Int. Ed. 52:1700–4
    [Google Scholar]
  168. 168.  Allpress CJ, Kleespies ST, Que L Jr 2015. Synthetic models of 2-oxoglutarate-dependent oxygenases. See Ref. 2 123–48
  169. 169.  Que L Jr, Tolman WB. 2008. Biologically inspired oxidation catalysis. Nature 455:333–40
    [Google Scholar]
/content/journals/10.1146/annurev-biochem-061516-044724
Loading
/content/journals/10.1146/annurev-biochem-061516-044724
Loading

Data & Media loading...

Supplemental Material

Supplementary Data

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error