763
Views
29
CrossRef citations to date
0
Altmetric
Review Article

The diverse role of TIGAR in cellular homeostasis and cancer

, , , &
Pages 1240-1249 | Received 26 Dec 2017, Accepted 09 Jun 2018, Published online: 04 Oct 2018

References

  • Hsu PP, Sabatini DM. Cancer cell metabolism: Warburg and beyond. Cell. 2008;134:703–707.
  • Schmitt CA, Fridman JS, Yang M, et al. A senescence program controlled by p53 and p16INK4a contributes to the outcome of cancer therapy. Cell. 2002;109:335–346.
  • Chipuk JE, Kuwana T, Bouchier-Hayes L, et al. Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis. Science. 2004;303:1010–1014.
  • Matoba S, Kang JG, Patino WD, et al. p53 regulates mitochondrial respiration. Science. 2006;312:1650–1653.
  • Crighton D, Wilkinson S, O’Prey J, et al. DRAM, a p53-induced modulator of autophagy, is critical for apoptosis. Cell. 2006;126:121–134.
  • Bensaad K, Tsuruta A, Selak MA, et al. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell. 2006;126:107–120.
  • Cheung EC, Ludwig RL, Vousden KH. Mitochondrial localization of TIGAR under hypoxia stimulates HK2 and lowers ROS and cell death. Proc Natl Acad Sci USA. 2012;109:20491–20496.
  • Wang CK, Ahmed MM, Jiang Q, et al. Melatonin ameliorates hypoglycemic stress-induced brain endothelial tight junction injury by inhibiting protein nitration of TP53-induced glycolysis and apoptosis regulator. J Pineal Res. 2017;63:e12440.
  • Li H, Jogl G. Structural and biochemical studies of TIGAR (TP53-induced glycolysis and apoptosis regulator). J Biol Chem. 2009;284:1748–1754.
  • Gerin I, Noël G, Bolsée J, et al. Identification of TP53-induced glycolysis and apoptosis regulator (TIGAR) as the phosphoglycolate-independent 2,3-bisphosphoglycerate phosphatase. Biochem J. 2014;458:439–448.
  • Yang J, Yan R, Roy A, et al. The I-TASSER Suite: protein structure and function prediction. Nat Methods. 2015;12:7–8.
  • Wang SJ, Li D, Ou Y, et al. Acetylation is crucial for p53-mediated ferroptosis and tumor suppression. Cell Rep. 2016;17:366–373.
  • Lee P, Hock AK, Vousden KH, Cheung EC. p53- and p73-independent activation of TIGAR expression in vivo. Cell Death Dis. 2015;6:e1842.
  • Li M, Sun M, Cao L, et al. A TIGAR-regulated metabolic pathway is critical for protection of brain ischemia. J Neurosci. 2014;34:7458–7471.
  • Zou S, Wang X, Deng L, et al. CREB, another culprit for TIGAR promoter activity and expression. Biochem Biophys Res Commun. 2013;439:481–486.
  • Zou S, Gu Z, Ni P, et al. SP1 plays a pivotal role for basal activity of TIGAR promoter in liver cancer cell lines. Mol Cell Biochem. 2012;359:17–23.
  • Rajendran R, Arva R, Ashour H, et al. Acetylation mediated by the p300/CBP-associated factor determines cellular energy metabolic pathways in cancer. Int J Oncol. 2013;42:1961–1972.
  • Xu X, Liu C, Bao J. Hypoxia-induced hsa-miR-101 promotes glycolysis by targeting TIGAR mRNA in clear cell renal cell carcinoma. Mol Med Rep. 2017;15:1373–1378.
  • Chen S, Li P, Li J, et al. MiR-144 inhibits proliferation and induces apoptosis and autophagy in lung cancer cells by targeting TIGAR. Cell Physiol Biochem. 2015;35:997–1007.
  • Cao L, Chen J, Li M, et al. Endogenous level of TIGAR in brain is associated with vulnerability of neurons to ischemic injury. Neurosci Bull. 2015;31:527–540.
  • Yu HP, Xie JM, Li B, et al. TIGAR regulates DNA damage and repair through pentosephosphate pathway and Cdk5-ATM pathway. Sci Rep. 2015;5:9853.
  • Rigden DJ. The histidine phosphatase superfamily: structure and function. Biochem J. 2008;409:333–348.
  • Monge L, Mojena M, Ortega JL, et al. Chlorpropamide raises fructose-2,6-bisphosphate concentration and inhibits gluconeogenesis in isolated rat hepatocytes. Diabetes. 1986;35:89–96.
  • Rodríguez-Gil JE, Gómez-Foix AM, Fillat C, et al. Activation by vanadate of glycolysis in hepatocytes from diabetic rats. Diabetes. 1991;40:1355–1359.
  • Choi IY, Wu C, Okar DA, et al. Elucidation of the role of fructose 2,6-bisphosphate in the regulation of glucose fluxes in mice using in vivo 13C NMR measurements of hepatic carbohydrate metabolism. FEBS J. 2002;269:4418–4426.
  • Wu C, Khan SA, Peng LJ, et al. Perturbation of glucose flux in the liver by decreasing F26P2 levels causes hepatic insulin resistance and hyperglycemia. Am J Physiol Endocrinol Metab. 2006;291:E536–E543.
  • Derdak Z, Lang CH, Villegas KA, et al. Activation of p53 enhances apoptosis and insulin resistance in a rat model of alcoholic liver disease. J Hepatol. 2011;54:164–172.
  • Qi Z, He J, Zhang Y, et al. Exercise training attenuates oxidative stress and decreases p53 protein content in skeletal muscle of type 2 diabetic Goto-Kakizaki rats. Free Radic Biol Med. 2011;50:794–800.
  • Sun M, Li M, Huang Q, et al. Ischemia/reperfusion-induced upregulation of TIGAR in brain is mediated by SP1 and modulated by ROS and hormones involved in glucose metabolism. Neurochem Int. 2015;80:99–109.
  • Zhou JH, Zhang TT, Song DD, et al. TIGAR contributes to ischemic tolerance induced by cerebral preconditioning through scavenging of reactive oxygen species and inhibition of apoptosis. Sci Rep. 2016;6:27096.
  • Jiang LB, Cao L, Ma YQ, et al. TIGAR mediates the inhibitory role of hypoxia on ROS production and apoptosis in rat nucleus pulposus cells. Osteoarthritis Cartilage. 2017;26:138–148.
  • Kim J, Devalaraja-Narashimha K, Padanilam BJ. TIGAR regulates glycolysis in ischemic kidney proximal tubules. Am J Physiol Renal Physiol. 2015;308:F298–F308.
  • Meng J, Lv Z, Qiao X, et al. The decay of Redox-stress Response Capacity is a substantive characteristic of aging: revising the redox theory of aging. Redox Biol. 2017;11:365–374.
  • Flinn LJ, Keatinge M, Bretaud S, et al. TigarB causes mitochondrial dysfunction and neuronal loss in PINK1 deficiency. Ann Neurol. 2013;74:837–847.
  • Hoshino A, Matoba S, Iwai-Kanai E, et al. p53-TIGAR axis attenuates mitophagy to exacerbate cardiac damage after ischemia. J Mol Cell Cardiol. 2012;52:175–184.
  • Kimata M, Matoba S, Iwai-Kanai E, et al. p53 and TIGAR regulate cardiac myocyte energy homeostasis under hypoxic stress. Am J Physiol Heart Circ Physiol. 2010; 299:H1908–H1916.
  • Martinez-Outschoorn UE, Trimmer C, Lin Z, et al. Autophagy in cancer associated fibroblasts promotes tumor cell survival: role of hypoxia, HIF1 induction and NFκB activation in the tumor stromal microenvironment. Cell Cycle. 2010;9:3515–3533.
  • Qian S, Li J, Hong M, et al. TIGAR cooperated with glycolysis to inhibit the apoptosis of leukemia cells and associated with poor prognosis in patients with cytogenetically normal acute myeloid leukemia. J Hematol Oncol. 2016;9:128.
  • Kumar B, Iqbal MA, Singh RK, et al. Resveratrol inhibits TIGAR to promote ROS induced apoptosis and autophagy. Biochimie. 2015;118:26–35.
  • Venkatanarayan A, Raulji P, Norton W, et al. Novel therapeutic interventions for p53-altered tumors through manipulation of its family members, p63 and p73. Cell Cycle. 2016;15:164–171.
  • Hong M, Xia Y, Zhu Y, et al. TP53-induced glycolysis and apoptosis regulator protects from spontaneous apoptosis and predicts poor prognosis in chronic lymphocytic leukemia. Leuk Res. 2016;50:72–77.
  • Hasegawa H, Yamada Y, Iha H, et al. Activation of p53 by Nutlin-3a, an antagonist of MDM2, induces apoptosis and cellular senescence in adult T-cell leukemia cells. Leukemia. 2009;23:2090–2101.
  • Wong E, Wong SC, Chan C, et al. TP53-induced glycolysis and apoptosis regulator promotes proliferation and invasiveness of nasopharyngeal carcinoma cells. Oncol Lett. 2015;9:569–574.
  • Zhao M, Fan J, Liu Y, et al. Oncogenic role of the TP53-induced glycolysis and apoptosis regulator in nasopharyngeal carcinoma through NF-κB pathway modulation. Int J Oncol. 2016;48:756–764.
  • Zhou X, Xie W, Li Q, et al. TIGAR is correlated with maximal standardized uptake value on FDG-PET and survival in non-small cell lung cancer. PLoS One. 2013;8:e80576.
  • Won KY, Lim SJ, Kim GY, et al. Regulatory role of p53 in cancer metabolism via SCO2 and TIGAR in human breast cancer. Hum Pathol. 2012;43:221–228.
  • Martinez-Outschoorn UE, Goldberg AF, Lin Z, et al. Anti-estrogen resistance in breast cancer is induced by the tumor microenvironment and can be overcome by inhibiting mitochondrial function in epithelial cancer cells. Cancer Biol Ther. 2011;12:924–938.
  • Lin CC, Cheng TL, Tsai WH, et al. Loss of the respiratory enzyme citrate synthase directly links the Warburg effect to tumor malignancy. Sci Rep. 2012;2:785.
  • Kim SH, Choi SI, Won KY, et al. Distinctive interrelation of p53 with SCO2, COX, and TIGAR in human gastric cancer. Pathol Res Pract. 2016;212:904–910.
  • Rajeshkumar NV, Dutta P, Yabuuchi S, et al. Therapeutic targeting of the Warburg effect in pancreatic cancer relies on an absence of p53 function. Cancer Res. 2015;75:3355–3364.
  • Dai Q, Yin Y, Liu W, et al. Two p53-related metabolic regulators, TIGAR and SCO2, contribute to oroxylin A-mediated glucose metabolism in human hepatoma HepG2 cells. Int J Biochem Cell Biol. 2013;45:1468–1478.
  • Liu S, Yan B, Lai W, et al. As a novel p53 direct target, bidirectional gene HspB2/αB-crystallin regulates the ROS level and Warburg effect. Biochim Biophys Acta. 2014;1839:592–603.
  • Al-Khayal K, Abdulla M, Al-Obeed O, et al. Identification of the TP53-induced glycolysis and apoptosis regulator in various stages of colorectal cancer patients. Oncol Rep. 2016;35:1281–1286.
  • Chesney J, Mitchell R, Benigni F, et al. An inducible gene product for 6-phosphofructo-2-kinase with an AU-rich instability element: role in tumor cell glycolysis and the Warburg effect. Proc Natl Acad Sci USA. 1999;96:3047–3052.
  • Ma R, Zhang W, Tang K, et al. Switch of glycolysis to gluconeogenesis by dexamethasone for treatment of hepatocarcinoma. Nat Commun. 2013;4:2508.
  • Ros S, Flöter J, Kaymak I, et al. 6-Phosphofructo-2-kinase/fructose-2,6-biphosphatase 4 is essential for p53-null cancer cells. Oncogene. 2017;36:3287–3299.
  • Dasgupta S, Rajapakshe K, Zhu B, et al. Metabolic enzyme PFKFB4 activates transcriptional coactivator SRC-3 to drive breast cancer. Nature. 2018;556:249–254.
  • Pavlides S, Whitaker-Menezes D, Castello-Cros R, et al. The reverse Warburg effect: aerobic glycolysis in cancer associated fibroblasts and the tumor stroma. Cell Cycle. 2009;8:3984–4001.
  • Ko YH, Domingo-Vidal M, Roche M, et al. TP53-inducible glycolysis and apoptosis regulator (TIGAR) metabolically reprograms carcinoma and stromal cells in breast cancer. J Biol Chem. 2016;291:26291–26303.
  • Shi XY, Xiong LX, Xiao L, et al. Downregulation of caveolin-1 upregulates the expression of growth factors and regulators in co-culture of fibroblasts with cancer cells. Mol Med Rep. 2016;13:744–752.
  • Wanka C, Steinbach JP, Rieger J. Tp53-induced glycolysis and apoptosis regulator (TIGAR) protects glioma cells from starvation-induced cell death by up-regulating respiration and improving cellular redox homeostasis. J Biol Chem. 2012;287:33436–33446.
  • Simon-Molas H, Calvo-Vidal MN, Castaño E, et al. Akt mediates TIGAR induction in HeLa cells following PFKFB3 inhibition. FEBS Lett. 2016;590:2915–2926.
  • Ghildiyal R, Sen E. CK2 induced RIG-I drives metabolic adaptations in IFNγ-treated glioma cells. Cytokine. 2017;89:219–228.
  • Cheung EC, Athineos D, Lee P, et al. TIGAR is required for efficient intestinal regeneration and tumorigenesis. Dev Cell. 2013;25:463–477.
  • Cheung EC, Lee P, Ceteci F, et al. Opposing effects of TIGAR- and RAC1-derived ROS on Wnt-driven proliferation in the mouse intestine. Genes Dev. 2016;30:52–63.
  • Qi Z, He Q, Ji L, et al. Antioxidant supplement inhibits skeletal muscle constitutive autophagy rather than fasting-induced autophagy in mice. Oxid Med Cell Longev. 2014;2014:315896.
  • Kim SJ, Jung HJ, Lim CJ. Reactive oxygen species-dependent down-regulation of tumor suppressor genes PTEN, USP28, DRAM, TIGAR, and CYLD under oxidative stress. Biochem Genet. 2013;51:901–915.
  • Bensaad K, Cheung EC, Vousden KH. Modulation of intracellular ROS levels by TIGAR controls autophagy. EMBO J. 2009;28:3015–3026.
  • Xie JM, Li B, Yu HP, et al. TIGAR has a dual role in cancer cell survival through regulating apoptosis and autophagy. Cancer Res. 2014;74:5127–5138.
  • Ye L, Zhao X, Lu J, et al. Knockdown of TIGAR by RNA interference induces apoptosis and autophagy in HepG2 hepatocellular carcinoma cells. Biochem Biophys Res Commun. 2013;437:300–306.
  • Ko YH, Lin Z, Flomenberg N, et al. Glutamine fuels a vicious cycle of autophagy in the tumor stroma and oxidative mitochondrial metabolism in epithelial cancer cells: implications for preventing chemotherapy resistance. Cancer Biol Ther. 2011;12:1085–1097.
  • Wang J, Duan Z, Nugent Z, et al. Reprogramming metabolism by histone methyltransferase NSD2 drives endocrine resistance via coordinated activation of pentose phosphate pathway enzymes. Cancer Lett. 2016;378:69–79.
  • Martinez-Marignac V, Toban N, Bazile M, et al. Resistance to Dasatinib in primary chronic lymphocytic leukemia lymphocytes involves AMPK-mediated energetic re-programming. Oncotarget. 2013;4:2550–2566.
  • Zhang H, Gu C, Yu J, et al. Radiosensitization of glioma cells by TP53-induced glycolysis and apoptosis regulator knockdown is dependent on thioredoxin-1 nuclear translocation. Free Radic Biol Med. 2014;69:239–248.
  • Zhang Y, Chen F, Tai G, et al. TIGAR knockdown radiosensitizes TrxR1-overexpressing glioma in vitro and in vivo via inhibiting Trx1 nuclear transport. Sci Rep. 2017;7:42928.
  • Yin L, Kosugi M, Kufe D. Inhibition of the MUC1-C oncoprotein induces multiple myeloma cell death by down-regulating TIGAR expression and depleting NADPH. Blood. 2012;119:810–816.
  • Nutt LK, Margolis SS, Jensen M, et al. Metabolic regulation of oocyte cell death through the CaMKII-mediated phosphorylation of caspase-2. Cell. 2005;123:89–103.
  • DeBerardinis RJ, Mancuso A, Daikhin E, et al. Beyond aerobic glycolysis: transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc Natl Acad Sci USA. 2007;104:19345–19350.
  • Yin L, Kufe T, Avigan D, et al. Targeting MUC1-C is synergistic with bortezomib in downregulating TIGAR and inducing ROS-mediated myeloma cell death. Blood. 2014;123:2997–3006.
  • Jain S, Stroopinsky D, Yin L, et al. Mucin 1 is a potential therapeutic target in cutaneous T-cell lymphoma. Blood. 2015;126:354–362.
  • Hasegawa M, Sinha RK, Kumar M, et al. Intracellular targeting of the oncogenic MUC1-C protein with a novel GO-203 nanoparticle formulation. Clin Cancer Res. 2015;21:2338–2347.
  • Ahmad R, Alam M, Hasegawa M, et al. Targeting MUC1-C inhibits the AKT-S6K1-elF4A pathway regulating TIGAR translation in colorectal cancer. Mol Cancer. 2017;16:33.
  • Peña-Rico MA, Calvo-Vidal MN, Villalonga-Planells R, et al. TP53 induced glycolysis and apoptosis regulator (TIGAR) knockdown results in radiosensitization of glioma cells. Radiother Oncol. 2011;101:132–139.
  • Iwao C, Shidoji Y. Upregulation of energy metabolism-related, p53-target TIGAR and SCO2 in HuH-7 cells with p53 mutation by geranylgeranoic acid treatment. Biomed Res. 2015;36:371–381.
  • Madan E, Gogna R, Kuppusamy P, et al. TIGAR induces p53-mediated cell-cycle arrest by regulation of RB-E2F1 complex. Br J Cancer. 2012;107:516–526.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.