1932

Abstract

The cytosolic selenoprotein thioredoxin reductase 1 (TrxR1, TXNRD1), and to some extent mitochondrial TrxR2 (TXNRD2), can be inhibited by a wide range of electrophilic compounds. Many such compounds also yield cytotoxicity toward cancer cells in culture or in mouse models, and most compounds are likely to irreversibly modify the easily accessible selenocysteine residue in TrxR1, thereby inhibiting its normal activity to reduce cytosolic thioredoxin (Trx1, TXN) and other substrates of the enzyme. This leads to an oxidative challenge. In some cases, the inhibited forms of TrxR1 are not catalytically inert and are instead converted to prooxidant NADPH oxidases, named SecTRAPs, thus further aggravating the oxidative stress, particularly in cells expressing higher levels of the enzyme. In this review, the possible molecular and cellular consequences of these effects are discussed in relation to cancer therapy, with a focus on outstanding questions that should be addressed if targeted TrxR1 inhibition is to be further developed for therapeutic use.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-pharmtox-052220-102509
2022-01-06
2024-04-24
Loading full text...

Full text loading...

/deliver/fulltext/pharmtox/62/1/annurev-pharmtox-052220-102509.html?itemId=/content/journals/10.1146/annurev-pharmtox-052220-102509&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    Purohit V, Simeone DM, Lyssiotis CA. 2019. Metabolic regulation of redox balance in cancer. Cancers 11:7955
    [Google Scholar]
  2. 2. 
    Sies H, Jones DP. 2020. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat. Rev. Mol. Cell Biol. 21:363–83
    [Google Scholar]
  3. 3. 
    Sies H, Berndt C, Jones DP. 2017. Oxidative stress. Annu. Rev. Biochem. 86:715–48
    [Google Scholar]
  4. 4. 
    Harris IS, Treloar AE, Inoue S, Sasaki M, Gorrini C et al. 2015. Glutathione and thioredoxin antioxidant pathways synergize to drive cancer initiation and progression. Cancer Cell 27:211–22
    [Google Scholar]
  5. 5. 
    Gorrini C, Harris IS, Mak TW 2013. Modulation of oxidative stress as an anticancer strategy. Nat. Rev. Drug Discov. 12:931–47
    [Google Scholar]
  6. 6. 
    Harris IS, DeNicola GM. 2020. The complex interplay between antioxidants and ROS in cancer. Trends Cell Biol 30:440–51
    [Google Scholar]
  7. 7. 
    Stafford WC, Peng X, Olofsson MH, Zhang X, Luci DK et al. 2018. Irreversible inhibition of cytosolic thioredoxin reductase 1 as a mechanistic basis for anticancer therapy. Sci. Transl. Med. 10:eaaf7444
    [Google Scholar]
  8. 8. 
    Anestal K, Prast-Nielsen S, Cenas N, Arner ES 2008. Cell death by SecTRAPs: thioredoxin reductase as a prooxidant killer of cells. PLOS ONE 3:e1846
    [Google Scholar]
  9. 9. 
    Arner ESJ. 2017. Targeting the selenoprotein thioredoxin reductase 1 for anticancer therapy. Adv. Cancer Res. 136:139–51
    [Google Scholar]
  10. 10. 
    Cebula M, Schmidt EE, Arnér ESJ. 2015. TrxR1 as a potent regulator of the Nrf2-Keap1 response system. Antioxid Redox. Signal. 23:823–53
    [Google Scholar]
  11. 11. 
    Fang J, Lu J, Holmgren A 2005. Thioredoxin reductase is irreversibly modified by curcumin: a novel molecular mechanism for its anticancer activity. J. Biol. Chem. 280:25284–90
    [Google Scholar]
  12. 12. 
    Becker K, Gromer S, Schirmer RH, Müller S. 2000. Thioredoxin reductase as a pathophysiological factor and drug target. Eur. J. Biochem. 267:6118–25
    [Google Scholar]
  13. 13. 
    Zhang B, Zhang J, Peng S, Liu R, Li X et al. 2016. Thioredoxin reductase inhibitors: a patent review. Expert Opin. Ther. Pat. 27:5547–56
    [Google Scholar]
  14. 14. 
    Wang L, Yang Z, Fu J, Yin H, Xiong K et al. 2012. Ethaselen: a potent mammalian thioredoxin reductase 1 inhibitor and novel organoselenium anticancer agent. Free Radic. Biol. Med. 52:898–908
    [Google Scholar]
  15. 15. 
    Watson J. 2013. Oxidants, antioxidants and the current incurability of metastatic cancers. Open Biol 3:120144
    [Google Scholar]
  16. 16. 
    Arnér ES. 2020. Perspectives of TrxR1-based cancer therapies. Oxidative Stress: Eustress and Distress H Sies 639–67 Cambridge, MA: Academic
    [Google Scholar]
  17. 17. 
    Mitsuishi Y, Motohashi H, Yamamoto M. 2012. The Keap1-Nrf2 system in cancers: stress response and anabolic metabolism. Front. Oncol. 2:200
    [Google Scholar]
  18. 18. 
    Bhatia M, McGrath KL, Di Trapani G, Charoentong P, Shah F et al. 2016. The thioredoxin system in breast cancer cell invasion and migration. Redox Biol 8:68–78
    [Google Scholar]
  19. 19. 
    Lincoln DT, Ali Emadi EM, Tonissen KF, Clarke FM 2003. The thioredoxin-thioredoxin reductase system: over-expression in human cancer. Anticancer Res 23:2425–33
    [Google Scholar]
  20. 20. 
    Xu J, Cheng Q, Arner ES. 2016. Details in the catalytic mechanism of mammalian thioredoxin reductase 1 revealed using point mutations and juglone-coupled enzyme activities. Free Radic. Biol. Med. 94:110–20
    [Google Scholar]
  21. 21. 
    Cheng Q, Antholine WE, Myers JM, Kalyanaraman B, Arnér ESJ, Myers CR 2010. The selenium-independent inherent pro-oxidant NADPH oxidase activity of mammalian thioredoxin reductase and its selenium-dependent direct peroxidase activities. J. Biol. Chem. 285:21708–23
    [Google Scholar]
  22. 22. 
    Anestål K, Prast-Nielsen S, Cenas N, Arnér ESJ 2008. Cell death by SecTRAPs—thioredoxin reductase as a prooxidant killer of cells. PLOS ONE 3:e1846
    [Google Scholar]
  23. 23. 
    Peng X, Xu J, Arner ES. 2012. Thiophosphate and selenite conversely modulate cell death induced by glutathione depletion or cisplatin: effects related to activity and Sec contents of thioredoxin reductase. Biochem. J. 447:167–74
    [Google Scholar]
  24. 24. 
    Eriksson SE, Prast-Nielsen S, Flaberg E, Szekely L, Arner ES 2009. High levels of thioredoxin reductase 1 modulate drug-specific cytotoxic efficacy. Free Radic. Biol. Med. 47:1661–71
    [Google Scholar]
  25. 25. 
    Busker S, Qian W, Haraldsson M, Espinosa B, Johansson L et al. 2020. Irreversible TrxR1 inhibitors block STAT3 activity and induce cancer cell death. Sci. Adv. 6:eaax7945
    [Google Scholar]
  26. 26. 
    Cheng Q, Sandalova T, Lindqvist Y, Arnér ESJ. 2009. Crystal structure and catalysis of the selenoprotein thioredoxin reductase 1. J. Biol. Chem. 284:3998–4008
    [Google Scholar]
  27. 27. 
    Zhong L, Arner ES, Holmgren A. 2000. Structure and mechanism of mammalian thioredoxin reductase: The active site is a redox-active selenolthiol/selenenylsulfide formed from the conserved cysteine-selenocysteine sequence. PNAS 97:5854–59
    [Google Scholar]
  28. 28. 
    Zhong L, Arnér ESJ, Ljung J, Åslund F, Holmgren A 1998. Rat and calf thioredoxin reductase are homologous to glutathione reductase with a carboxyl-terminal elongation containing a conserved catalytically active penultimate selenocysteine residue. J. Biol. Chem. 273:8581–91
    [Google Scholar]
  29. 29. 
    Peng X, Mandal PK, Kaminskyy VO, Lindqvist A, Conrad M, Arner ES 2014. Sec-containing TrxR1 is essential for self-sufficiency of cells by control of glucose-derived H2O2. Cell Death Dis 5:e1235
    [Google Scholar]
  30. 30. 
    Dagnell M, Schmidt EE, Arner ESJ. 2018. The A to Z of modulated cell patterning by mammalian thioredoxin reductases. Free Radic. Biol. Med. 115:484–96
    [Google Scholar]
  31. 31. 
    Lei XG, Zhu JH, Cheng WH, Bao Y, Ho YS et al. 2016. Paradoxical roles of antioxidant enzymes: basic mechanisms and health implications. Physiol. Rev. 96:307–64
    [Google Scholar]
  32. 32. 
    Iverson SV, Eriksson S, Xu J, Prigge JR, Talago EA et al. 2013. A Txnrd1-dependent metabolic switch alters hepatic lipogenesis, glycogen storage, and detoxification. Free Radic. Biol. Med. 63:369–80
    [Google Scholar]
  33. 33. 
    Fernandes AP, Holmgren A. 2004. Glutaredoxins: glutathione-dependent redox enzymes with functions far beyond a simple thioredoxin backup system. Antioxid. Redox Signal. 6:63–74
    [Google Scholar]
  34. 34. 
    Lu J, Chew E-H, Holmgren A. 2007. Targeting thioredoxin reductase is a basis for cancer therapy by arsenic trioxide. PNAS 104:12288
    [Google Scholar]
  35. 35. 
    Holmgren A. 1989. Thioredoxin and glutaredoxin systems. J. Biol. Chem. 264:13963–66
    [Google Scholar]
  36. 36. 
    Murata H, Ihara Y, Nakamura H, Yodoi J, Sumikawa K, Kondo T. 2003. Glutaredoxin exerts an antiapoptotic effect by regulating the redox state of Akt. J. Biol. Chem. 278:50226–33
    [Google Scholar]
  37. 37. 
    Eriksson S, Prigge JR, Talago EA, Arner ES, Schmidt EE 2015. Dietary methionine can sustain cytosolic redox homeostasis in the mouse liver. Nat. Commun. 6:6479
    [Google Scholar]
  38. 38. 
    Prigge JR, Eriksson S, Iverson SV, Meade TA, Capecchi MR et al. 2012. Hepatocyte DNA replication in growing liver requires either glutathione or a single allele of txnrd1. Free Radic. Biol. Med. 52:803–10
    [Google Scholar]
  39. 39. 
    Mandal PK, Schneider M, Kolle P, Kuhlencordt P, Forster H et al. 2010. Loss of thioredoxin reductase 1 renders tumors highly susceptible to pharmacologic glutathione deprivation. Cancer Res 70:9505–14
    [Google Scholar]
  40. 40. 
    Hudemann C, Lonn ME, Godoy JR, Zahedi Avval F, Capani F et al. 2009. Identification, expression pattern, and characterization of mouse glutaredoxin 2 isoforms. Antioxid. Redox Signal. 11:1–14
    [Google Scholar]
  41. 41. 
    Lonn ME, Hudemann C, Berndt C, Cherkasov V, Capani F et al. 2008. Expression pattern of human glutaredoxin 2 isoforms: identification and characterization of two testis/cancer cell-specific isoforms. Antioxid. Redox Signal. 10:547–57
    [Google Scholar]
  42. 42. 
    Du Y, Zhang H, Lu J, Holmgren A 2012. Glutathione and glutaredoxin act as a backup of human thioredoxin reductase 1 to reduce thioredoxin 1 preventing cell death by aurothioglucose. J. Biol. Chem. 287:38210–19
    [Google Scholar]
  43. 43. 
    Du Y, Zhang H, Zhang X, Lu J, Holmgren A 2013. Thioredoxin 1 is inactivated due to oxidation induced by peroxiredoxin under oxidative stress and reactivated by the glutaredoxin system. J. Biol. Chem. 288:32241–47
    [Google Scholar]
  44. 44. 
    Branco V, Coppo L, Solá S, Lu J, Rodrigues CMP et al. 2017. Impaired cross-talk between the thioredoxin and glutathione systems is related to ASK-1 mediated apoptosis in neuronal cells exposed to mercury. Redox Biol 13:278–87
    [Google Scholar]
  45. 45. 
    Zhang H, Du Y, Zhang X, Lu J, Holmgren A 2013. Glutaredoxin 2 reduces both thioredoxin 2 and thioredoxin 1 and protects cells from apoptosis induced by auranofin and 4-hydroxynonenal. Antioxid. Redox Signal. 21:669–81
    [Google Scholar]
  46. 46. 
    Johansson C, Lillig CH, Holmgren A. 2004. Human mitochondrial glutaredoxin reduces S-glutathionylated proteins with high affinity accepting electrons from either glutathione or thioredoxin reductase. J. Biol. Chem. 279:7537–43
    [Google Scholar]
  47. 47. 
    Espinosa B, Arnér ESJ. 2019. Thioredoxin-related protein of 14 kDa as a modulator of redox signalling pathways. Br. J. Pharmacol. 176:544–53
    [Google Scholar]
  48. 48. 
    Pader I, Sengupta R, Cebula M, Xu J, Lundberg JO et al. 2014. Thioredoxin-related protein of 14 kDa is an efficient L-cystine reductase and S-denitrosylase. PNAS 111:6964–69
    [Google Scholar]
  49. 49. 
    Jeong W, Chang TS, Boja ES, Fales HM, Rhee SG. 2004. Roles of TRP14, a thioredoxin-related protein in tumor necrosis factor-α signaling pathways. J. Biol. Chem. 279:3151–59
    [Google Scholar]
  50. 50. 
    Jeong W, Yoon HW, Lee SR, Rhee SG 2004. Identification and characterization of TRP14, a thioredoxin-related protein of 14 kDa. New insights into the specificity of thioredoxin function. J. Biol. Chem. 279:3142–50
    [Google Scholar]
  51. 51. 
    Dagnell M, Frijhoff J, Pader I, Augsten M, Boivin B et al. 2013. Selective activation of oxidized PTP1B by the thioredoxin system modulates PDGF-β receptor tyrosine kinase signaling. PNAS 110:13398–403
    [Google Scholar]
  52. 52. 
    Cox AG, Winterbourn CC, Hampton MB. 2009. Mitochondrial peroxiredoxin involvement in antioxidant defence and redox signalling. Biochem. J. 425:313–25
    [Google Scholar]
  53. 53. 
    Rabilloud T, Heller M, Rigobello MP, Bindoli A, Aebersold R, Lunardi J 2001. The mitochondrial antioxidant defence system and its response to oxidative stress. Proteomics 1:1105–10
    [Google Scholar]
  54. 54. 
    Hanschmann EM, Lonn ME, Schutte LD, Funke M, Godoy JR et al. 2010. Both thioredoxin 2 and glutaredoxin 2 contribute to the reduction of the mitochondrial 2-Cys peroxiredoxin Prx3. J. Biol. Chem. 285:40699–705
    [Google Scholar]
  55. 55. 
    Gencheva R, Cheng Q, Arner ESJ. 2018. Efficient selenocysteine-dependent reduction of toxoflavin by mammalian thioredoxin reductase. Biochim. Biophys. Acta Gen. Subj. 1862:2511–17
    [Google Scholar]
  56. 56. 
    Rackham O, Shearwood A-MJ, Thyer R, McNamara E, Davies SMK et al. 2011. Substrate and inhibitor specificities differ between human cytosolic and mitochondrial thioredoxin reductases: implications for development of specific inhibitors. Free Radic. Biol. Med. 50:689–99
    [Google Scholar]
  57. 57. 
    Scalcon V, Bindoli A, Rigobello MP. 2018. Significance of the mitochondrial thioredoxin reductase in cancer cells: an update on role, targets and inhibitors. Free Radic. Biol. Med. 127:62–79
    [Google Scholar]
  58. 58. 
    Cox AG, Brown KK, Arner ES, Hampton MB 2008. The thioredoxin reductase inhibitor auranofin triggers apoptosis through a Bax/Bak-dependent process that involves peroxiredoxin 3 oxidation. Biochem. Pharmacol. 76:1097–109
    [Google Scholar]
  59. 59. 
    Brown KK, Eriksson SE, Arner ES, Hampton MB 2008. Mitochondrial peroxiredoxin 3 is rapidly oxidized in cells treated with isothiocyanates. Free Radic. Biol. Med. 45:494–502
    [Google Scholar]
  60. 60. 
    Talbot S, Nelson R, Self WT 2008. Arsenic trioxide and auranofin inhibit selenoprotein synthesis: implications for chemotherapy for acute promyelocytic leukaemia. Br. J. Pharmacol. 154:940–48
    [Google Scholar]
  61. 61. 
    Krishnamurthy D, Karver MR, Fiorillo E, Orru V, Stanford SM et al. 2008. Gold(I)-mediated inhibition of protein tyrosine phosphatases: a detailed in vitro and cellular study. J. Med. Chem. 51:4790–95
    [Google Scholar]
  62. 62. 
    Miller WH Jr., Schipper HM, Lee JS, Singer J, Waxman S. 2002. Mechanisms of action of arsenic trioxide. Cancer Res 62:3893–903
    [Google Scholar]
  63. 63. 
    Cenas N, Nivinskas H, Anusevicius Z, Sarlauskas J, Lederer F, Arner ES. 2004. Interactions of quinones with thioredoxin reductase: a challenge to the antioxidant role of the mammalian selenoprotein. J. Biol. Chem. 279:2583–92
    [Google Scholar]
  64. 64. 
    Rigobello MP, Folda A, Baldoin MC, Scutari G, Bindoli A. 2005. Effect of auranofin on the mitochondrial generation of hydrogen peroxide. Role of thioredoxin reductase. Free Radic. Res. 39:687–95
    [Google Scholar]
  65. 65. 
    Inarrea P, Moini H, Han D, Rettori D, Aguilo I et al. 2007. Mitochondrial respiratory chain and thioredoxin reductase regulate intermembrane Cu,Zn-superoxide dismutase activity: implications for mitochondrial energy metabolism and apoptosis. Biochem. J. 405:173–79
    [Google Scholar]
  66. 66. 
    Rackham O, Shearwood AM, Thyer R, McNamara E, Davies SM et al. 2011. Substrate and inhibitor specificities differ between human cytosolic and mitochondrial thioredoxin reductases: implications for development of specific inhibitors. Free Radic. Biol. Med. 50:689–99
    [Google Scholar]
  67. 67. 
    Reddy CA, Somepalli V, Golakoti T, Kanugula AK, Karnewar S et al. 2014. Mitochondrial-targeted curcuminoids: a strategy to enhance bioavailability and anticancer efficacy of curcumin. PLOS ONE 9:e89351
    [Google Scholar]
  68. 68. 
    Jayakumar S, Patwardhan RS, Pal D, Singh B, Sharma D et al. 2017. Mitochondrial targeted curcumin exhibits anticancer effects through disruption of mitochondrial redox and modulation of TrxR2 activity. Free Radic. Biol. Med. 113:530–38
    [Google Scholar]
  69. 69. 
    Liang B, Shao W, Zhu C, Wen G, Yue X et al. 2016. Mitochondria-targeted approach: remarkably enhanced cellular bioactivities of TPP2a as selective inhibitor and probe toward TrxR. ACS Chem. Biol. 11:425–34
    [Google Scholar]
  70. 70. 
    Lu J, Chew EH, Holmgren A. 2007. Targeting thioredoxin reductase is a basis for cancer therapy by arsenic trioxide. PNAS 104:12288–93
    [Google Scholar]
  71. 71. 
    Ganan-Gomez I, Wei Y, Yang H, Boyano-Adanez MC, Garcia-Manero G. 2013. Oncogenic functions of the transcription factor Nrf2. Free Radic. Biol. Med. 65:750–64
    [Google Scholar]
  72. 72. 
    Higgins LG, Hayes JD. 2011. The cap'n'collar transcription factor Nrf2 mediates both intrinsic resistance to environmental stressors and an adaptive response elicited by chemopreventive agents that determines susceptibility to electrophilic xenobiotics. Chem. Biol. Interact. 192:37–45
    [Google Scholar]
  73. 73. 
    Bai X, Chen Y, Hou X, Huang M, Jin J. 2016. Emerging role of NRF2 in chemoresistance by regulating drug-metabolizing enzymes and efflux transporters. Drug Metab. Rev. 48:541–67
    [Google Scholar]
  74. 74. 
    Baird L, Yamamoto M. 2020. The molecular mechanisms regulating the KEAP1-NRF2 pathway. Mol. Cell. Biol. 40:13e00099-20
    [Google Scholar]
  75. 75. 
    Probst BL, McCauley L, Trevino I, Wigley WC, Ferguson DA. 2015. Cancer cell growth is differentially affected by constitutive activation of NRF2 by KEAP1 deletion and pharmacological activation of NRF2 by the synthetic triterpenoid, RTA 405. PLOS ONE 10:e0135257
    [Google Scholar]
  76. 76. 
    Gorrini C, Baniasadi PS, Harris IS, Silvester J, Inoue S et al. 2013. BRCA1 interacts with Nrf2 to regulate antioxidant signaling and cell survival. J. Exp. Med. 210:1529–44
    [Google Scholar]
  77. 77. 
    Taguchi K, Yamamoto M. 2020. The KEAP1-NRF2 system as a molecular target of cancer treatment. Cancers 13:46
    [Google Scholar]
  78. 78. 
    Fabrizio FP, Sparaneo A, Muscarella LA. 2020. NRF2 regulation by noncoding RNAs in cancers: the present knowledge and the way forward. Cancers 12:3621
    [Google Scholar]
  79. 79. 
    Rundlöf A-K, Carlsten M, Arnér ESJ 2001. The core promoter of human thioredoxin reductase 1: Cloning, transcriptional activity and Oct-1, Sp1 and Sp3 binding reveal a housekeeping-type promoter for the AU-rich element-regulated gene. J. Biol. Chem. 276:30542–51
    [Google Scholar]
  80. 80. 
    Hintze KJ, Wald KA, Zeng H, Jeffery EH, Finley JW 2003. Thioredoxin reductase in human hepatoma cells is transcriptionally regulated by sulforaphane and other electrophiles via an antioxidant response element. J. Nutr. 133:2721–27
    [Google Scholar]
  81. 81. 
    Cebula M, Schmidt EE, Arner ES. 2015. TrxR1 as a potent regulator of the Nrf2-Keap1 response system. Antioxid. Redox Signal. 23:823–53
    [Google Scholar]
  82. 82. 
    Busker S, Qian W, Haraldsson M, Espinosa B, Johansson L et al. 2020. Irreversible TrxR1 inhibitors block STAT3 activity and induce cancer cell death. Sci. Adv. 6:eaax7945
    [Google Scholar]
  83. 83. 
    Demaria M, Camporeale A, Poli V 2014. STAT3 and metabolism: how many ways to use a single molecule?. Int. J. Cancer 135:1997–2003
    [Google Scholar]
  84. 84. 
    Lee H, Pal SK, Reckamp K, Figlin RA, Yu H. 2011. STAT3: a target to enhance antitumor immune response. Curr. Top. Microbiol. Immunol. 344:41–59
    [Google Scholar]
  85. 85. 
    Lee HJ, Zhuang G, Cao Y, Du P, Kim HJ, Settleman J 2014. Drug resistance via feedback activation of Stat3 in oncogene-addicted cancer cells. Cancer Cell 26:207–21
    [Google Scholar]
  86. 86. 
    Busker S, Page B, Arnér ES 2020. To inhibit TrxR1 is to inactivate STAT3—inhibition of TrxR1 enzymatic function by STAT3 small molecule inhibitors. Redox Biol 36:101646
    [Google Scholar]
  87. 87. 
    Sobotta MC, Liou W, Stocker S, Talwar D, Oehler M et al. 2015. Peroxiredoxin-2 and STAT3 form a redox relay for H2O2 signaling. Nat. Chem. Biol. 11:64–70
    [Google Scholar]
  88. 88. 
    Bykov VJ, Issaeva N, Shilov A, Hultcrantz M, Pugacheva E et al. 2002. Restoration of the tumor suppressor function to mutant p53 by a low-molecular-weight compound. Nat. Med. 8:282–88
    [Google Scholar]
  89. 89. 
    Peng X, Zhang MQ, Conserva F, Hosny G, Selivanova G et al. 2013. APR-246/PRIMA-1MET inhibits thioredoxin reductase 1 and converts the enzyme to a dedicated NADPH oxidase. Cell Death Dis 4:e881
    [Google Scholar]
  90. 90. 
    Mohell N, Alfredsson J, Fransson Å, Uustalu M, Byström S et al. 2015. APR-246 overcomes resistance to cisplatin and doxorubicin in ovarian cancer cells. Cell Death Dis 6:e1794
    [Google Scholar]
  91. 91. 
    Tessoulin B, Descamps G, Moreau P, Maiga S, Lode L et al. 2014. PRIMA-1Met induces myeloma cell death independent of p53 by impairing the GSH/ROS balance. Blood 124:1626–36
    [Google Scholar]
  92. 92. 
    Haffo L, Lu J, Bykov VJN, Martin SS, Ren X et al. 2018. Inhibition of the glutaredoxin and thioredoxin systems and ribonucleotide reductase by mutant p53-targeting compound APR-246. Sci. Rep. 8:12671
    [Google Scholar]
  93. 93. 
    Sachweh MC, Stafford WC, Drummond CJ, McCarthy AR, Higgins M et al. 2015. Redox effects and cytotoxic profiles of MJ25 and auranofin towards malignant melanoma cells. Oncotarget 6:16488–506
    [Google Scholar]
  94. 94. 
    Dabiri Y, Abu El Maaty MA, Chan HY, Wölker J, Ott I et al. 2019. p53-Dependent anti-proliferative and pro-apoptotic effects of a Gold(I) N-heterocyclic carbene (NHC) complex in colorectal cancer cells. Front. Oncol. 9:438
    [Google Scholar]
  95. 95. 
    Shi Y, Nikulenkov F, Zawacka-Pankau J, Li H, Gabdoulline R et al. 2014. ROS-dependent activation of JNK converts p53 into an efficient inhibitor of oncogenes leading to robust apoptosis. Cell Death Differ 21:612–23
    [Google Scholar]
  96. 96. 
    Singh M, Zhou X, Chen X, Santos GS, Peuget S et al. 2020. Identification and targeting of selective vulnerability rendered by tamoxifen resistance. Breast Cancer Res 22:80
    [Google Scholar]
  97. 97. 
    Cassidy PB, Edes K, Nelson CC, Parsawar K, Fitzpatrick FA, Moos PJ. 2006. Thioredoxin reductase is required for the inactivation of tumor suppressor p53 and for apoptosis induced by endogenous electrophiles. Carcinogenesis 27:2538–49
    [Google Scholar]
  98. 98. 
    Moos PJ, Edes K, Cassidy P, Massuda E, Fitzpatrick FA 2003. Electrophilic prostaglandins and lipid aldehydes repress redox-sensitive transcription factors p53 and hypoxia-inducible factor by impairing the selenoprotein thioredoxin reductase. J. Biol. Chem. 278:745–50
    [Google Scholar]
  99. 99. 
    Hedström E, Eriksson S, Zawacka-Pankau J, Arnér ES, Selivanova G 2009. p53-Dependent inhibition of TrxR1 contributes to the tumor-specific induction of apoptosis by RITA. Cell Cycle 8:3584–91
    [Google Scholar]
  100. 100. 
    Xu J, Eriksson SE, Cebula M, Sandalova T, Hedstrom E et al. 2015. The conserved Trp114 residue of thioredoxin reductase 1 has a redox sensor-like function triggering oligomerization and crosslinking upon oxidative stress related to cell death. Cell Death Dis 6:e1616
    [Google Scholar]
  101. 101. 
    Peyser BD, Hermone A, Salamoun JM, Burnett JC, Hollingshead MG et al. 2019. Specific RITA modification produces hyperselective cytotoxicity while maintaining in vivo antitumor efficacy. Mol. Cancer Ther. 18:1765–74
    [Google Scholar]
  102. 102. 
    Ueno M, Masutani H, Arai RJ, Yamauchi A, Hirota K et al. 1999. Thioredoxin-dependent redox regulation of p53-mediated p21 activation. J. Biol. Chem. 274:35809–15
    [Google Scholar]
  103. 103. 
    Hanson S, Kim E, Deppert W 2005. Redox factor 1 (Ref-1) enhances specific DNA binding of p53 by promoting p53 tetramerization. Oncogene 24:1641–47
    [Google Scholar]
  104. 104. 
    Seemann S, Hainaut P. 2005. Roles of thioredoxin reductase 1 and APE/Ref-1 in the control of basal p53 stability and activity. Oncogene 24:3853–63
    [Google Scholar]
  105. 105. 
    Liu B, Chen Y, St. Clair DK 2008. ROS and p53: a versatile partnership. Free Radic. Biol. Med. 44:1529–35
    [Google Scholar]
  106. 106. 
    Arnér ESJ, Holmgren A. 2006. The thioredoxin system in cancer. Semin. Cancer Biol. 16:420–26
    [Google Scholar]
  107. 107. 
    Lillig CH, Holmgren A. 2007. Thioredoxin and related molecules—from biology to health and disease. Antioxid. Redox Signal. 9:25–47
    [Google Scholar]
  108. 108. 
    Zahedi Avval F, Holmgren A 2009. Molecular mechanisms of thioredoxin and glutaredoxin as hydrogen donors for mammalian S phase ribonucleotide reductase. J. Biol. Chem. 284:8233–40
    [Google Scholar]
  109. 109. 
    Haffo L, Lu J, Bykov VJN, Martin SS, Ren X et al. 2018. Inhibition of the glutaredoxin and thioredoxin systems and ribonucleotide reductase by mutant p53-targeting compound APR-246. Sci. Rep. 8:12671
    [Google Scholar]
  110. 110. 
    Hashemy SI, Ungerstedt JS, Zahedi Avval F, Holmgren A 2006. Motexafin gadolinium, a tumor-selective drug targeting thioredoxin reductase and ribonucleotide reductase. J. Biol. Chem. 281:10691–97
    [Google Scholar]
  111. 111. 
    Muri J, Heer S, Matsushita M, Pohlmeier L, Tortola L et al. 2018. The thioredoxin-1 system is essential for fueling DNA synthesis during T-cell metabolic reprogramming and proliferation. Nat. Commun. 9:1851
    [Google Scholar]
  112. 112. 
    Wakasugi N, Tagaya Y, Wakasugi H, Mitsui A, Maeda M et al. 1990. Adult T-cell leukemia-derived factor/thioredoxin, produced by both human T-lymphotropic virus type I- and Epstein-Barr virus-transformed lymphocytes, acts as an autocrine growth factor and synergizes with interleukin 1 and interleukin 2. PNAS 87:8282–86
    [Google Scholar]
  113. 113. 
    Tagaya Y, Maeda Y, Mitsui A, Kondo N, Matsui H et al. 1989. ATL-derived factor (ADF), an IL-2 receptor/Tac inducer homologous to thioredoxin; possible involvement of dithiol-reduction in the IL-2 receptor induction. EMBO J 8:757–64
    [Google Scholar]
  114. 114. 
    Backman E, Bergh AC, Lagerdahl I, Rydberg B, Sundstrom C et al. 2007. Thioredoxin, produced by stromal cells retrieved from the lymph node microenvironment, rescues chronic lymphocytic leukemia cells from apoptosis in vitro. Haematologica 92:1495–504
    [Google Scholar]
  115. 115. 
    Nakamura H, De Rosa SC, Yodoi J, Holmgren A, Ghezzi P et al. 2001. Chronic elevation of plasma thioredoxin: inhibition of chemotaxis and curtailment of life expectancy in AIDS. PNAS 98:2688–93
    [Google Scholar]
  116. 116. 
    Schenk H, Vogt M, Droge W, Schulze-Osthof K. 1996. Thioredoxin as a potent costimulus of cytokine expression. J. Immunol. 156:765–71
    [Google Scholar]
  117. 117. 
    Pekkari K, Goodarzi MT, Scheynius A, Holmgren A, Avila-Cariño J 2005. Truncated thioredoxin (Trx80) induces differentiation of human CD14+ monocytes into a novel cell type (TAMs) via activation of the MAP kinases p38, ERK, and JNK. Blood 105:1598–605
    [Google Scholar]
  118. 118. 
    Pekkari K, Holmgren A. 2004. Truncated thioredoxin: physiological functions and mechanism. Antioxid. Redox Signal. 6:53–61
    [Google Scholar]
  119. 119. 
    Mahmood DF, Abderrazak A, Couchie D, Lunov O, Diderot V et al. 2013. Truncated thioredoxin (Trx-80) promotes pro-inflammatory macrophages of the M1 phenotype and enhances atherosclerosis. J. Cell Physiol. 228:1577–83
    [Google Scholar]
  120. 120. 
    Nakamura H, Masutani H, Yodoi J. 2006. Extracellular thioredoxin and thioredoxin-binding protein 2 in control of cancer. Semin. Cancer Biol. 16:444–51
    [Google Scholar]
  121. 121. 
    Nordberg J, Arnér ESJ. 2001. Reactive oxygen species, antioxidants, and the mammalian thioredoxin system. Free Radic. Biol. Med. 31:1287–312
    [Google Scholar]
  122. 122. 
    Peng W, Zhou Z, Zhong Y, Sun Y, Wang Y et al. 2019. Plasma activity of thioredoxin reductase as a novel biomarker in gastric cancer. Sci. Rep. 9:19084
    [Google Scholar]
  123. 123. 
    Ye S, Chen X, Yao Y, Li Y, Sun R et al. 2019. Thioredoxin reductase as a novel and efficient plasma biomarker for the detection of non-small cell lung cancer: a large-scale, multicenter study. Sci. Rep. 9:2652
    [Google Scholar]
  124. 124. 
    Zhang L, Cheng Q, Zhang L, Wang Y, Merrill GF et al. 2016. Serum thioredoxin reductase is highly increased in mice with hepatocellular carcinoma and its activity is restrained by several mechanisms. Free Radic. Biol. Med. 99:426–35
    [Google Scholar]
  125. 125. 
    Delgado-Buenrostro NL, Medina-Reyes EI, Lastres-Becker I, Freyre-Fonseca V, Ji Z et al. 2015. Nrf2 protects the lung against inflammation induced by titanium dioxide nanoparticles: a positive regulator role of Nrf2 on cytokine release. Environ. Toxicol. 30:782–92
    [Google Scholar]
  126. 126. 
    Oshi M, Angarita FA, Tokumaru Y, Yan L, Matsuyama R et al. 2020. High expression of NRF2 is associated with increased tumor-infiltrating lymphocytes and cancer immunity in ER-positive/HER2-negative breast cancer. Cancers 12:3856
    [Google Scholar]
  127. 127. 
    Kong H, Chandel NS. 2018. Regulation of redox balance in cancer and T cells. J. Biol. Chem. 293:7499–507
    [Google Scholar]
  128. 128. 
    Holmdahl R, Sareila O, Pizzolla A, Winter S, Hagert C et al. 2013. Hydrogen peroxide as an immunological transmitter regulating autoreactive T cells. Antioxid. Redox Signal. 18:1463–74
    [Google Scholar]
  129. 129. 
    Devadas S, Zaritskaya L, Rhee SG, Oberley L, Williams MS 2002. Discrete generation of superoxide and hydrogen peroxide by T cell receptor stimulation: selective regulation of mitogen-activated protein kinase activation and Fas ligand expression. J. Exp. Med. 195:59–70
    [Google Scholar]
  130. 130. 
    Laniewski NG, Grayson JM. 2004. Antioxidant treatment reduces expansion and contraction of antigen-specific CD8+ T cells during primary but not secondary viral infection. J. Virol. 78:11246–57
    [Google Scholar]
  131. 131. 
    Wang R, Dillon CP, Shi LZ, Milasta S, Carter R et al. 2011. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity 35:871–82
    [Google Scholar]
  132. 132. 
    Previte DM, O'Connor EC, Novak EA, Martins CP, Mollen KP, Piganelli JD. 2017. Reactive oxygen species are required for driving efficient and sustained aerobic glycolysis during CD4+ T cell activation. PLOS ONE 12:e0175549
    [Google Scholar]
  133. 133. 
    Franchina DG, Dostert C, Brenner D. 2018. Reactive oxygen species: involvement in T cell signaling and metabolism. Trends Immunol 39:489–502
    [Google Scholar]
  134. 134. 
    Angelini G, Gardella S, Ardy M, Ciriolo MR, Filomeni G et al. 2002. Antigen-presenting dendritic cells provide the reducing extracellular microenvironment required for T lymphocyte activation. PNAS 99:1491–96
    [Google Scholar]
  135. 135. 
    Mak TW, Grusdat M, Duncan GS, Dostert C, Nonnenmacher Y et al. 2017. Glutathione primes T cell metabolism for inflammation. Immunity 46:675–89
    [Google Scholar]
  136. 136. 
    Chakraborty P, Chatterjee S, Kesarwani P, Thyagarajan K, Iamsawat S et al. 2019. Thioredoxin-1 improves the immunometabolic phenotype of antitumor T cells. J. Biol. Chem. 294:9198–212
    [Google Scholar]
  137. 137. 
    Deleted in proof
  138. 138. 
    Morzadec C, Macoch M, Sparfel L, Kerdine-Römer S, Fardel O, Vernhet L 2014. Nrf2 expression and activity in human T lymphocytes: stimulation by T cell receptor activation and priming by inorganic arsenic and tert-butylhydroquinone. Free Radic. Biol. Med. 71:133–45
    [Google Scholar]
  139. 139. 
    Zagorski JW, Turley AE, Freeborn RA, VanDenBerg KR, Dover HE et al. 2018. Differential effects of the Nrf2 activators tBHQ and CDDO-Im on the early events of T cell activation. Biochem. Pharmacol. 147:67–76
    [Google Scholar]
  140. 140. 
    Rashida Gnanaprakasam JN, Wu R, Wang R 2018. Metabolic reprogramming in modulating T cell reactive oxygen species generation and antioxidant capacity. Front. Immunol. 9:1075
    [Google Scholar]
  141. 141. 
    Mao Y, Poschke I, Kiessling R 2014. Tumour-induced immune suppression: role of inflammatory mediators released by myelomonocytic cells. J. Intern. Med. 276:154–70
    [Google Scholar]
  142. 142. 
    Cemerski S, Cantagrel A, Van Meerwijk JP, Romagnoli P. 2002. Reactive oxygen species differentially affect T cell receptor-signaling pathways. J. Biol. Chem. 277:19585–93
    [Google Scholar]
  143. 143. 
    Mougiakakos D, Johansson CC, Kiessling R. 2009. Naturally occurring regulatory T cells show reduced sensitivity toward oxidative stress-induced cell death. Blood 113:3542–45
    [Google Scholar]
  144. 144. 
    Ando T, Mimura K, Johansson CC, Hanson MG, Mougiakakos D et al. 2008. Transduction with the antioxidant enzyme catalase protects human T cells against oxidative stress. J. Immunol. 181:8382–90
    [Google Scholar]
  145. 145. 
    Lee HL, Jang JW, Lee SW, Yoo SH, Kwon JH et al. 2019. Inflammatory cytokines and change of Th1/Th2 balance as prognostic indicators for hepatocellular carcinoma in patients treated with transarterial chemoembolization. Sci. Rep. 9:3260
    [Google Scholar]
  146. 146. 
    Wang X, Dong H, Li Q, Li Y, Hong A 2015. Thioredoxin induces Tregs to generate an immunotolerant tumor microenvironment in metastatic melanoma. Oncoimmunology 4:e1027471
    [Google Scholar]
  147. 147. 
    Mougiakakos D, Johansson CC, Jitschin R, Böttcher M, Kiessling R. 2011. Increased thioredoxin-1 production in human naturally occurring regulatory T cells confers enhanced tolerance to oxidative stress. Blood 117:857–61
    [Google Scholar]
  148. 148. 
    Yu H, Pardoll D, Jove R 2009. STATs in cancer inflammation and immunity: a leading role for STAT3. Nat. Rev. Cancer 9:798–809
    [Google Scholar]
  149. 149. 
    Lee H, Jeong AJ, Ye SK. 2019. Highlighted STAT3 as a potential drug target for cancer therapy. BMB Rep 52:415–23
    [Google Scholar]
  150. 150. 
    Kitamura H, Ohno Y, Toyoshima Y, Ohtake J, Homma S et al. 2017. Interleukin-6/STAT3 signaling as a promising target to improve the efficacy of cancer immunotherapy. Cancer Sci 108:1947–52
    [Google Scholar]
  151. 151. 
    Furtek SL, Backos DS, Matheson CJ, Reigan P. 2016. Strategies and approaches of targeting STAT3 for cancer treatment. ACS Chem. Biol. 11:308–18
    [Google Scholar]
  152. 152. 
    Ghosh S, Mukherjee S, Choudhury S, Gupta P, Adhikary A et al. 2015. Reactive oxygen species in the tumor niche triggers altered activation of macrophages and immunosuppression: role of fluoxetine. Cell. Signal. 27:1398–412
    [Google Scholar]
  153. 153. 
    Ruffell B, Coussens LM. 2015. Macrophages and therapeutic resistance in cancer. Cancer Cell 27:462–72
    [Google Scholar]
  154. 154. 
    Santi A, Caselli A, Ranaldi F, Paoli P, Mugnaioni C et al. 2015. Cancer associated fibroblasts transfer lipids and proteins to cancer cells through cargo vesicles supporting tumor growth. Biochim. Biophys. Acta Mol. Cell Res. 1853:3211–23
    [Google Scholar]
  155. 155. 
    Chai D, Zhang L, Xi S, Cheng Y, Jiang H, Hu R. 2018. Nrf2 activation induced by Sirt1 ameliorates acute lung injury after intestinal ischemia/reperfusion through NOX4-mediated gene regulation. Cell. Physiol. Biochem. 46:781–92
    [Google Scholar]
  156. 156. 
    Brigelius-Flohé R. 2008. Selenium compounds and selenoproteins in cancer. Chem. Biodivers. 5:389–95
    [Google Scholar]
  157. 157. 
    Locy ML, Rogers LK, Prigge JR, Schmidt EE, Arner ES, Tipple TE 2012. Thioredoxin reductase inhibition elicits Nrf2-mediated responses in Clara cells: implications for oxidant-induced lung injury. Antioxid. Redox Signal. 17:1407–16
    [Google Scholar]
  158. 158. 
    Rollins MF, van der Heide DM, Weisend CM, Kundert JA, Comstock KM et al. 2010. Hepatocytes lacking thioredoxin reductase 1 have normal replicative potential during development and regeneration. J. Cell Sci. 123:2402–12
    [Google Scholar]
  159. 159. 
    Smith AD, Guidry CA, Morris VC, Levander OA 1999. Aurothioglucose inhibits murine thioredoxin reductase activity in vivo. J. Nutr. 129:194–98
    [Google Scholar]
/content/journals/10.1146/annurev-pharmtox-052220-102509
Loading
/content/journals/10.1146/annurev-pharmtox-052220-102509
Loading

Data & Media loading...

  • 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