Review article Oncometabolites: Unconventional triggers of oncogenic signalling cascades Marco Sciacovelli, Christian Frezza n Medical Research Council Cancer Unit, University of Cambridge, Hutchison/MRC Research Centre, Box 197, Cambridge Biomedical Campus, Cambridge CB2 0XZ, United Kingdom article info Article history: Received 31 January 2016 Received in revised form 11 April 2016 Accepted 19 April 2016 Available online 23 April 2016 Keywords: Mitochondria Cancer Oncometabolites FH SDH IDH Fumarate Succinate 2-Hydroxyglutarate abstract Cancer is a complex and heterogeneous disease thought to be caused by multiple genetic lesions. The recent finding that enzymes of the tricarboxylic acid (TCA) cycle are mutated in cancer rekindled the hypothesis that altered metabolism might also have a role in cellular transformation. Attempts to link mitochondrial dysfunction to cancer uncovered the unexpected role of small molecule metabolites, now known as oncometabolites, in tumorigenesis. In this review, we describe how oncometabolites can contribute to tumorigenesis. We propose that lesions of oncogenes and tumour suppressors are only one of the possible routes to tumorigenesis, which include accumulation of oncometabolites triggered by environmental cues. & 2016 The Authors. Published by Elsevier Inc. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). 1. Background Cancer is a complex and multifactorial disease. Although its malignant features have been known for centuries, it was not until the advent of modern biology that the molecular determinants of cancer transformation have been elucidated. In 1911, pioneering work from Peyton Rous showed that avian sarcomas were transmissible to other healthy fowls through a filtrate of the tumours devoid of cells [1,2]. Later on, the agent present in those extracts and responsible for tumour formation was identified as a retrovirus, later called Rous Sarcoma Virus. This important discovery started the field of tumour virology and led to the identification of the oncogene v-Src [3–6], first in retroviruses and then in normal avian DNA [7]. The emerging idea was that cancer was caused by alterations of the genome. Since then, other oncogenes, including MYC, RAS, ERBB, PI3K [2,6] and the first tumour suppressor gene RB1 [8,9] were discovered. The causative role of RB1 inactivation in retinoblastoma formation reinforced the concept of cancer initiation driven by genomic alterations. These discoveries led Knudson and colleagues to hypothesise a “multiple-hit” model of tumorigenesis, where multiple genetic alterations are required to achieve full blown transformation [8]. We now know that tumorigenesis requires the acquisition of multiple enabling features, the hallmarks of cancer, among which metabolic rewiring is becoming increasingly recognised [10,11]. In a seminal paper, Shim et al. showed that Lactate Dehydrogenase A (LDHA) is a target of the proto-oncogene MYC and is required for MYC-induced anchorage-independent growth in both human and mouse cellular models [12]. Since then, scientists have uncovered several aspects of the metabolic reprogramming of cancer, and realised that not only dysregulated metabolism is required to sustain proliferation but it also affects tumour microenvironment and the immune response [11]. The discovery that mutations in the metabolic genes Fumarate hydratase (FH) [13], Succinate dehydrogenase (SDH) [14–17], and Isocitrate dehydrogenase (IDH) [18–21] lead to cancer further supported a primary role of metabolic alterations in tumorigenesis. Thanks to these discoveries, a novel paradigm is emerging whereby mitochondrial metabolites that accumulate in these conditions act as oncogenic signalling molecules, becoming bona fide oncometabolites. Recent data suggest that reprogramming of cellular metabolism occurs both as direct and indirect consequence of oncogenic mutations and that environmental cues, such as hypoxia, could affect the metabolic phenotype of cancer cells and the abundance of oncometabolites, amplifying oncogenic cascades. In this review we describe the main oncogenic functions Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/freeradbiomed Free Radical Biology and Medicine http://dx.doi.org/10.1016/j.freeradbiomed.2016.04.025 0891-5849/& 2016 The Authors. Published by Elsevier Inc. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). n Corresponding author. E-mail address: cf366@MRC-CU.cam.ac.uk (C. Frezza). Free Radical Biology and Medicine 100 (2016) 175–181 of oncometabolites and how their abundance can be affected by genetic mutations and environmental cues. 2. TCA cycle enzymes mutations and the emerging paradigm of oncometabolite-driven tumorigenesis SDH was the first mitochondrial enzyme found mutated in cancer [14]. It was the first time that mutations of a mitochondrial enzyme, once thought to be incompatible with life [22], were linked to tumour predisposition. This and the subsequent discovery of FH mutations in renal cancer catalysed a substantial effort to elucidate the molecular links between mitochondrial dysfunction and tumorigenesis. These major findings are reported below. 2.1. Succinate dehydrogenase (SDH) SDH is an enzyme of the TCA cycle involved in the conversion of succinate to fumarate and a key component of the mitochondrial respiratory chain. This enzyme is composed of four subunits, SDHA, SDHB, SDHC, SDHD; and two assembly factors, SDHF1 and SDHF2 [23]. Mutations in SDH are found in familial paraganglioma and pheochromocytoma [14–17,24,25], renal carcinomas [26], T-Cell leukaemia [27], and gastrointestinal stromal tumours [28]. SDH deficiency causes profound metabolic changes. Recent work showed that mouse SDH-deficient cells have a high demand of extracellular pyruvate and utilise glucose-derived carbons for aspartate biosynthesis through pyruvate carboxylation [29,30]. Interestingly, this metabolic rewiring could be used to selectively target Sdhb-/- cells [29]. One of the most striking features of SDH-deficient cells is the accumulation of succinate [31], a metabolite implicated in tumorigenesis and, for this reasons, recently defined an oncometabolite [32]. Among its many functions, succinate is a competitive inhibitor of α-ketoglutarate (aKG)-dependent dioxygenases (aKGDD), a class of enzymes involved in a plethora of biological processes. For instance, succinate inhibits prolyl-hydroxylases (PHDs), aKGDDs involved in the degradation of Hypoxia Inducible Factor (HIF), leading to the aberrant stabilisation of HIFs even when oxygen is abundant, a condition called pseudohypoxia [33]. Succinate inhibits other aKGDDs, including Ten-Eleven Translocation proteins (TETs), enzymes involved in DNA demethylation [34,35], leading to CpG island hypermethylation [36]. Succinate also causes the inhibition of Histone Lysine Demethylases (KDMs), aKGDDs involved in histone demethylation [34], causing even further epigenetic changes [36,37]. Interestingly, DNA hypermethylation phenotype was associated with dedifferentiation and increased invasion potential of SDH-deficient tumours [36,38]. However, the molecular mechanisms behind this phenotypic switch are still under investigation. 2.2. Fumarate hydratase FH is an enzyme of the TCA cycle that converts fumarate to malate. Whilst homozygous FH mutations cause fumaric aciduria, a condition associated with infantile encephalopathy and brain malformations [39], heterozygous FH mutations followed by the loss of heterozygosity of the second allele cause Hereditary Leiomyomatosis and Renal Cell Cancer (HLRCC) [13,40]. FH is also mutated in paraganglioma, pheochromocytoma [41,42], downregulated in sporadic clear cell carcinomas [43] and deleted in neuroblastoma [44]. Cristal structure of human FH showed that clinically-relevant mutations affect evolutionary conserved regions involved in either the catalytic activity or the folding and stability of the protein [45], leading to abnormal accumulation of fumarate [46–48]. Loss of FH also leads to a complex rewiring of cell metabolism. For instance, FH loss leads to an increased uptake of glutamine that is diverted into haem synthesis and bilirubin excretion to maintain mitochondrial NADH production and mitochondrial potential. Also, the accumulation of fumarate in FHdeficient cells leads to the reversal of the urea cycle enzyme argininosuccinate lyase (ASL), causing the production of argininosuccinate from fumarate and arginine [47,48]. This metabolic rewiring makes FH-deficient cells auxotrophic for arginine and sensitive to arginine-depriving agents such as arginine deiminase [48]. Interestingly, arginine depletion has been proposed as therapeutic intervention for renal cell carcinoma [49]. However, in this case, arginine auxotrophy is caused by inactivation of argininosuccinate synthase (ASS1), the urea cycle enzyme that converts aspartate and citrulline to argininosuccinate, which is then converted to arginine and fumarate by ASL. Inactivation of ASS1 has been shown to allow the diversion of aspartate from the urea cycle to pyrimidine biosynthesis, favouring tumour growth [50]. It is tempting to speculate that in FH-deficient cells the activity of ASS1 is inhibited by the accumulation of argininosuccinate, leading to increased pyrimidine biosynthesis. Consistent with an increased utilisation of aspartate, the intracellular levels of this metabolite are very low in FH-deficient cells [48]. However, the rate of pyrimidine synthesis is these cells has not been assessed yet. Fumarate has been implicated in tumorigenesis of HLRCC and, for this reason, included in the list of oncometabolites [51]. Similarly to succinate, fumarate inhibits several aKGDDs, including PHDs, leading to pseudohypoxia [52]. Of note, the non-canonical activation of NF-kB signalling by fumarate also contributes to the pseudohypoxic phenotype in FH-deficient cells [53]. Although pseudohypoxia has been considered an important driver of tumorigenesis, recent data showed that HIFs are dispensable for the formation of benign pre-tumorigenic lesions in Fh1-deficient mice [54]. Therefore, other mechanisms have been proposed to explain fumarate-dependent tumorigenesis. For instance, recent findings identified the Abelson murine leukemia viral oncogene homolog 1 (ABL-1) as potential driver in fumarate-dependent tumorigenesis [55]. Interestingly, ABL-1 inhibitors suppress the invasion properties of FH-deficient cells both in vitro and in vivo. Mechanistically, through ABL-1 activation, fumarate stimulates an antioxidant response mediated by transcription factor NFE2-related factor 2 (NRF2), and a metabolic rewiring through activation of mammalian target of rapamycin (mTOR)-HIF axis [55]. Fumarate was also shown to promote broad epigenetic changes [56], caused by inhibition of histone and DNA demethylases [34,35]. However, the relevance of these epigenetic changes in tumorigenesis is still under investigation. Finally, fumarate, a mild electrophilic molecule, was demonstrated to react with thiol residues of proteins through a process called succination [57–60]. This post-translational modification is a distinctive feature of FH-deficient tumours and now used for diagnostic purposes [58]. Although several succinated proteins have been identified in FH-deficient cells [60,61], the biological roles of this process are still under investigation. It was recently reported that succination of mitochondrial Aconitase (ACO2) impairs its enzymatic activity [61] and succination of Kelch-Like ECH-associated protein-1 (KEAP-1) inhibits its negative modulatory effect on the transcription factor NRF2 [54,62]. Although NRF2 activation has been previously reported as pro-tumorigenic event [63], its role in FH-deficient tumours is still debated. The thiol residue of the antioxidant tripeptide Glutathione (GSH) is also subject of succination [64,65] and its depletion causes oxidative stress, which induces senescence in primary FH-deficient kidney cells. Of note, genetic ablation of p21, a major player in senescence induction, induces transformation of benign cysts in FH-deficient mice indicating the tumour-suppressive role of senescence in FH-mediated 176 M. Sciacovelli, C. Frezza / Free Radical Biology and Medicine 100 (2016) 175–181 tumorigenesis [65]. Besides revealing potential mechanisms of tumorigenesis caused by loss of FH, these works led to the identification of fumarate as important regulator of redox homeostasis, which goes beyond FH deficiency. For instance, Jin et al. showed that fumarate generated within the TCA cycle can regulate the activity of the antioxidant protein Glutathione Peroxidase (GPx), modulating the antioxidant capacity of the cell. Of note, this function of fumarate is not caused by succination but, rather, by the binding of fumarate to the protein [66]. 2.3. Isocitrate dehydrogenase (IDH1/IDH2) Isocitrate dehydrogenases (IDHs) are homodimeric enzymes responsible for the reversible oxidative decarboxylation of isocitrate to aKG. Three different isoforms of IDH have been described, which have distinct subcellular compartmentalisation. IDH3 is a NADþ-dependent mitochondrial enzyme, core component of the TCA cycle. The other two IDH isoforms (IDH1 and IDH2) are NADPþ-dependent proteins expressed respectively in cytosol and mitochondria [25,67]. Heterozygous missense mutations affecting IDH1 and IDH2 were found in gliomas [18,19] and in acute myeloid leukaemia (AML) [20,21]. At odds with SDH and FH, IDH mutations are gain-of-function mutations and confer to the enzyme the ability to produce the oncometabolite D-2-hydroxyglutarate (D-2HG) [68,69]. 2HG, the reduced form of aKG, is naturally present in two optic isomers,D-2HG and L-2HG. In normal cells, 2HG is a minor by-product of metabolism and its levels are kept low by the activity of a conserved family of proteins, the L/D-2-hydroxyglutarate dehydrogenases (L2HGDH and D2HGDH) [70,71], which convert 2HG to aKG. Homozygous germline mutations in these two enzymes are responsible for a severe form of 2HG aciduria characterised by developmental abnormalities and premature death in children [71,72]. Moreover, mutations in L2HGDH have been associated with brain tumours [73] and, recently, with kidney cancer [74]. Both isomers of 2HG can affect the enzymatic activity of aKGDD. Whilst the role of D-2HG in PHD inhibition and HIF stabilisation is still controversial [75], this metabolite was shown to inhibit both TETs and KDMs [76,77]. Consistently, IDH1-mutant tumours exhibit DNA hypermethylation in gliomas [78] and leukemia [79], and histone hypermethylation [80]. Evidence that TET2 and IDH1/2 mutations are mutually exclusive in AML tumours [79] supports the notion that TETs inhibition and the ensuing DNA hypermethylation are instrumental to IDH-driven tumorigenesis. This hypothesis has been recently corroborated by the finding that DNA hypermethylation causes reduced CCCTC-binding factor (CTCF) binding to DNA, leading to aberrant activation of the oncogene Platelet-derived growth factor receptor (PDGFRA) [81] in gliomas. Other mechanisms have been proposed to contribute to 2HG-dependent tumorigenesis in IDHmutant tumours. For instance, 2HG accumulation inhibits ATP Synthase within the mitochondria activating a series of downstream signals that involve mTOR suppression [82]. Regardless the molecular underpinnings, research on IDH-mutant tumours uncovered an unexpected but powerful role of 2HG as oncogenic molecule. In support of this role, it was demonstrated that the incubation of cells with D-2HG promotes cytokine independence and alters differentiation in hematopoietic cells, proving compelling evidence that a small molecule is sufficient to transform cells [83]. Interestingly, the role of 2HG in tumorigenesis goes beyond IDH-mutant tumours. For instance, 2HG accumulation can be driven by a metabolic reprogramming caused by MYC activation in aggressive breast cancer, leading to DNA hypermethylation akin to that observed in IDH mutant tumours [84]. Also, 2HG can be produced by the promiscuous activity of the enzyme Phosphoglycerate Dehydrogenase, another important oncogene in breast cancer [85]. 3. Succinate, fumarate, and 2HG: overlapping and distinct functions The brief description above suggests that signalling cascades elicited by fumarate, succinate, and 2HG, share some common targets but also exhibit distinct features, which can explain the Fig. 1. Schematic representation of overlapping features of tumours harbouring FH, SDH, and IDH mutations. Mutations in FH, SDH, and IDH lead to the accumulation of fumarate, succinate, and 2HG, respectively, activating a plethora of signalling cascades. Converging signatures, mainly mediated by aKGDD inhibition, include pseudohypoxia, histone and DNA hypermethylation. The colour of text indicates metabolic alterations (red), epigenetic alterations (blue), post-translational modifications (green) and other pro-tumorigenic alterations (black) elicited by these metabolites. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) M. Sciacovelli, C. Frezza / Free Radical Biology and Medicine 100 (2016) 175–181 177 different types of tumours associated with their accumulation (Fig. 1). The converging signatures, mostly mediated by aKGDD inhibition, include pseudohypoxia and broad epigenetic changes, such as DNA and histones hypermethylation. Interestingly, recent studies showed that pseudohypoxic genes, besides being regulated by HIFs, are also transcriptionally controlled by TETs. Therefore, epigenetic changes are required for a full pseudohypoxic response triggered by these metabolites. However, these metabolites have different IC50 for TETs and KDMs [35], suggesting that their epigenetic effect might have different outcomes, with fumarate being the most effective TETs inhibitor and 2HG the poorest. These metabolites exert other distinct biological functions. As indicated above, fumarate accumulation leads to protein succination, triggering a plethora of biological changes that may synergise with or counteract aKGDDs inhibition. Also, accumulation of 2HG has been shown to increase protein succinylation via inhibition of SDH and subsequent accumulation of succinyl-CoA [86]. This post translation modification leads to a reprogramming of mitochondrial function and induces resistance to apoptosis. Finally, fumarate and 2HG elicit opposite effects on mTOR signalling, with important consequences for the development of specific tumour types [55,82]. Therefore, although characterised by a very similar chemical structure, succinate, fumarate, and 2HG, appear to have distinct biological roles, well beyond inhibition of aKGDDs. 4. Environmental cues regulates oncometabolite production The observation that 2HG is sufficient to promote tumorigenesis [83] raised the possibility that environmental cues that increase this metabolite could contribute to tumorigenesis, without underpinning IDH mutations. In support to this hypothesis, it has been shown that hypoxia leads to the production of 2HG (Fig. 2), either via reductive carboxylation [87] or via promiscuous substrate usage of LDHA [88] and Malic Dehydrogenase 1/2 [89]. Since hypoxia is a common feature of solid tumours [90] it is possible that hypoxia-driven production of 2HG elicits (epi)genetic changes that drive or amplify the process of tumorigenesis. 2-HG is not the only oncometabolite whose levels are altered by hypoxia. During ischemia, succinate significantly accumulates in multiple organs and its oxidation is responsible for ROS generation during reperfusion [91]. Succinate was also shown to increase under hypoxic conditions, in cancer cells cultured in 3D scaffolds [92], consistent with its role as hypoxia sensing molecule. Other cues trigger oncometabolite accumulation. For instance, fumarate was shown to accumulate in hyperglycemic conditions [59,93] (Fig. 2), likely as a consequence of mitochondrial dysfunction caused by glucose accumulation [94,95]. We hypothesise that fumarate accumulation observed in diabetes [93,95], could, at least in part, explain the increase cancer risk in these patients. Together, these studies seem to suggest that environmental or nutritional cues may cause dysregulation of mitochondrial function, leading to oncometabolite accumulation in the absence of underpinning oncogenic mutations. 5. Future perspectives Oncometabolites are emerging as key components of the communication between mitochondria and the nucleus. Chronic accumulation of these small molecules triggered by genetic or environmental cues may alter the epigenetic landscape of the cell eliciting oncogenic signalling cascades. This new paradigm of tumorigenesis challenges the role of gene mutations as the exclusive driving mechanism of tumorigenesis and suggests that the latter may be only one of the possible mechanisms that lead to transformation. Competing interests The authors declare no competing interests. Authors’ contribution MS and CF jointly wrote the manuscript. Authors’ information MS is a Research Associate in the laboratory of CF. CF is a group leader at the MRC Cancer Unit, University of Cambridge, Fig. 2. An oncometabolic perspective of tumorigenesis. Schematic representation of how oncometabolite affect the process of tumorigenesis. The indicated oncometabolites can accumulate as a consequence of mutations of TCA cycle enzymes or environmental cues, such as hypoxia or hyperglycemia. These metabolites can act as proper oncogenic triggers and can drive transformation even in the absence of genetic alterations. 178 M. Sciacovelli, C. Frezza / Free Radical Biology and Medicine 100 (2016) 175–181 Cambridge, UK. MS and CF are funded by an MRC Core Funding to the MRC Cancer Unit. References [1] P. Rous, A sarcoma of the fowl transmissible by an agent separable from the tumor cells, J. Exp. Med. 13 (4) (1911) 397–411. [2] R.A. Weiss, P.K. Vogt, 100 years of Rous sarcoma virus, J. Exp. Med. 208 (12) (2011) 2351–2355. [3] H.M. Temin, H. Rubin, Characteristics of an assay for Rous sarcoma virus and Rous sarcoma cells in tissue culture, Virology 6 (3) (1958) 669–688. [4] P.H. Duesberg, P.K. Vogt, Differences between the ribonucleic acids of transforming and nontransforming avian Tumor viruses, Proc. Natl. Acad. Sci. USA 67 (4) (1970) 1673–1680. [5] G.S. Martin, Rous sarcoma virus: a function required for the maintenance of the transformed state, Nature 227 (5262) (1970) 1021–1023. [6] P.K. Vogt, Retroviral oncogenes: a historical primer, Nat. Rev. Cancer 12 (9) (2012) 639–648. [7] D. Stehelin, H.E. Varmus, J.M. Bishop, P.K. Vogt, DNA related to the transforming gene(s) of avian sarcoma viruses is present in normal avian DNA, Nature 260 (5547) (1976) 170–173. [8] A.G. Knudson Jr., Mutation and cancer: statistical study of retinoblastoma, Proc. Natl. Acad. Sci. USA 68 (4) (1971) 820–823. [9] D.L. Burkhart, J. Sage, Cellular mechanisms of tumour suppression by the retinoblastoma gene, Nat. Rev. Cancer 8 (9) (2008) 671–682. [10] D. Hanahan, R.A. Weinberg, Hallmarks of cancer: the next generation, Cell 144 (5) (2011) 646–674. [11] N.N. Pavlova, C.B. Thompson, The emerging hallmarks of cancer metabolism, Cell Metab. 23 (1) (2016) 27–47. [12] H. Shim, C. Dolde, B.C. Lewis, C.S. Wu, G. Dang, R.A. Jungmann, R. Dalla-Favera, C.V. Dang, c-Myc transactivation of LDH-A: implications for tumor metabolism and growth, Proc. Natl. Acad. Sci. USA 94 (13) (1997) 6658–6663. [13] I.P. Tomlinson, N.A. Alam, A.J. Rowan, E. Barclay, E.E. Jaeger, D. Kelsell, I. Leigh, P. Gorman, H. Lamlum, S. Rahman, R.R. Roylance, S. Olpin, S. Bevan, K. Barker, N. Hearle, R.S. Houlston, M. Kiuru, R. Lehtonen, A. Karhu, S. Vilkki, P. Laiho, C. Eklund, O. Vierimaa, K. Aittomaki, M. Hietala, P. Sistonen, A. Paetau, R. Salovaara, R. Herva, V. Launonen, L.A. Aaltonen, Germline mutations in FH predispose to dominantly inherited uterine fibroids, skin leiomyomata and papillary renal cell cancer, Nat. Genet. 30 (4) (2002) 406–410. [14] B.E. Baysal, R.E. Ferrell, J.E. Willett-Brozick, E.C. Lawrence, D. Myssiorek, A. Bosch, A. van der Mey, P.E. Taschner, W.S. Rubinstein, E.N. Myers, C. W. Richard 3rd, C.J. Cornelisse, P. Devilee, B. Devlin, Mutations in SDHD, a mitochondrial complex II gene, in hereditary paraganglioma, Science 287 (5454) (2000) 848–851. [15] S. Niemann, U. Muller, Mutations in SDHC cause autosomal dominant paraganglioma, type 3, Nat. Genet. 26 (3) (2000) 268–270. [16] D. Astuti, F. Latif, A. Dallol, P.L. Dahia, F. Douglas, E. George, F. Skoldberg, E. S. Husebye, C. Eng, E.R. Maher, Gene mutations in the succinate dehydrogenase subunit SDHB cause susceptibility to familial pheochromocytoma and to familial paraganglioma, Am. J. Hum. Genet. 69 (1) (2001) 49–54. [17] H.X. Hao, O. Khalimonchuk, M. Schraders, N. Dephoure, J.P. Bayley, H. Kunst, P. Devilee, C.W. Cremers, J.D. Schiffman, B.G. Bentz, S.P. Gygi, D.R. Winge, H. Kremer, J. Rutter, SDH5, a gene required for flavination of succinate dehydrogenase, is mutated in paraganglioma, Science 325 (5944) (2009) 1139–1142. [18] D.W. Parsons, S. Jones, X. Zhang, J.C. Lin, R.J. Leary, P. Angenendt, P. Mankoo, H. Carter, I.M. Siu, G.L. Gallia, A. Olivi, R. McLendon, B.A. Rasheed, S. Keir, T. Nikolskaya, Y. Nikolsky, D.A. Busam, H. Tekleab, L.A. Diaz Jr., J. Hartigan, D. R. Smith, R.L. Strausberg, S.K. Marie, S.M. Shinjo, H. Yan, G.J. Riggins, D. D. Bigner, R. Karchin, N. Papadopoulos, G. Parmigiani, B. Vogelstein, V. E. Velculescu, K.W. Kinzler, An integrated genomic analysis of human glioblastoma multiforme, Science 321 (5897) (2008) 1807–1812. [19] H. Yan, D.W. Parsons, G. Jin, R. McLendon, B.A. Rasheed, W. Yuan, I. Kos, I. Batinic-Haberle, S. Jones, G.J. Riggins, H. Friedman, A. Friedman, D. Reardon, J. Herndon, K.W. Kinzler, V.E. Velculescu, B. Vogelstein, D.D. Bigner, IDH1 and IDH2 mutations in gliomas, N. Engl. J. Med. 360 (8) (2009) 765–773. [20] A. Green, P. Beer, Somatic mutations of IDH1 and IDH2 in the leukemic transformation of myeloproliferative neoplasms, N. Engl. J. Med. 362 (4) (2010) 369–370. [21] E.R. Mardis, L. Ding, D.J. Dooling, D.E. Larson, M.D. McLellan, K. Chen, D. C. Koboldt, R.S. Fulton, K.D. Delehaunty, S.D. McGrath, L.A. Fulton, D.P. Locke, V. J. Magrini, R.M. Abbott, T.L. Vickery, J.S. Reed, J.S. Robinson, T. Wylie, S. M. Smith, L. Carmichael, J.M. Eldred, C.C. Harris, J. Walker, J.B. Peck, F. Du, A. F. Dukes, G.E. Sanderson, A.M. Brummett, E. Clark, J.F. McMichael, R.J. Meyer, J. K. Schindler, C.S. Pohl, J.W. Wallis, X. Shi, L. Lin, H. Schmidt, Y. Tang, C. Haipek, M.E. Wiechert, J.V. Ivy, J. Kalicki, G. Elliott, R.E. Ries, J.E. Payton, P. Westervelt, M.H. Tomasson, M.A. Watson, J. Baty, S. Heath, W.D. Shannon, R. Nagarajan, D. C. Link, M.J. Walter, T.A. Graubert, J.F. DiPersio, R.K. Wilson, T.J. Ley, Recurring mutations found by sequencing an acute myeloid leukemia genome, N. Engl. J. Med., 361, (2009) 1058–1066. [22] P. Rustin, T. Bourgeron, B. Parfait, D. Chretien, A. Munnich, A. Rotig, Inborn errors of the Krebs cycle: a group of unusual mitochondrial diseases in human, Biochim. Biophys. Acta 1361 (2) (1997) 185–197. [23] J.G. Van Vranken, U. Na, D.R. Winge, J. Rutter, Protein-mediated assembly of succinate dehydrogenase and its cofactors, Crit. Rev. Biochem. Mol. Biol. 50 (2) (2015) 168–180. [24] C. Bardella, P.J. Pollard, I. Tomlinson, SDH mutations in cancer, Biochim. Biophys. Acta 1807 (11) (2011) 1432–1443. [25] D.C. Wallace, Mitochondria and cancer, Nat. Rev. Cancer 12 (10) (2012) 685–698. [26] S. Vanharanta, M. Buchta, S.R. McWhinney, S.K. Virta, M. Peczkowska, C. D. Morrison, R. Lehtonen, A. Januszewicz, H. Jarvinen, M. Juhola, J.P. Mecklin, E. Pukkala, R. Herva, M. Kiuru, N.N. Nupponen, L.A. Aaltonen, H.P. Neumann, C. Eng, Early-onset renal cell carcinoma as a novel extraparaganglial component of SDHB-associated heritable paraganglioma, Am. J. Hum. Genet. 74 (1) (2004) 153–159. [27] B.E. Baysal, A recurrent stop-codon mutation in succinate dehydrogenase subunit B gene in normal peripheral blood and childhood T-cell acute leukemia, PLoS One 2 (5) (2007) e436. [28] C.A. Stratakis, J.A. Carney, The triad of paragangliomas, gastric stromal tumours and pulmonary chondromas (Carney triad), and the dyad of paragangliomas and gastric stromal sarcomas (Carney-Stratakis syndrome): molecular genetics and clinical implications, J. Intern. Med. 266 (1) (2009) 43–52. [29] S. Cardaci, L. Zheng, G. MacKay, N.J. van den Broek, E.D. MacKenzie, C. Nixon, D. Stevenson, S. Tumanov, V. Bulusu, J.J. Kamphorst, A. Vazquez, S. Fleming, F. Schiavi, G. Kalna, K. Blyth, D. Strathdee, E. Gottlieb, Pyruvate carboxylation enables growth of SDH-deficient cells by supporting aspartate biosynthesis, Nat. Cell Biol. 17 (10) (2015) 1317–1326. [30] C. Lussey-Lepoutre, K.E. Hollinshead, C. Ludwig, M. Menara, A. Morin, L. J. Castro-Vega, S.J. Parker, M. Janin, C. Martinelli, C. Ottolenghi, C. Metallo, A. P. Gimenez-Roqueplo, J. Favier, D.A. Tennant, Loss of succinate dehydrogenase activity results in dependency on pyruvate carboxylation for cellular anabolism, Nat. Commun. 6 (2015) 8784. [31] P.J. Pollard, J.J. Briere, N.A. Alam, J. Barwell, E. Barclay, N.C. Wortham, T. Hunt, M. Mitchell, S. Olpin, S.J. Moat, I.P. Hargreaves, S.J. Heales, Y.L. Chung, J. R. Griffiths, A. Dalgleish, J.A. McGrath, M.J. Gleeson, S.V. Hodgson, R. Poulsom, P. Rustin, I.P. Tomlinson, Accumulation of Krebs cycle intermediates and overexpression of HIF1alpha in tumours which result from germline FH and SDH mutations, Hum. Mol. Genet. 14 (15) (2005) 2231–2239. [32] M. Yang, P.J. Pollard, Succinate: a new epigenetic hacker, Cancer Cell 23 (6) (2013) 709–711. [33] M.A. Selak, S.M. Armour, E.D. MacKenzie, H. Boulahbel, D.G. Watson, K. D. Mansfield, Y. Pan, M.C. Simon, C.B. Thompson, E. Gottlieb, Succinate links TCA cycle dysfunction to oncogenesis by inhibiting HIF-alpha prolyl hydroxylase, Cancer Cell 7 (1) (2005) 77–85. [34] M. Xiao, H. Yang, W. Xu, S. Ma, H. Lin, H. Zhu, L. Liu, Y. Liu, C. Yang, Y. Xu, S. Zhao, D. Ye, Y. Xiong, K.L. Guan, Inhibition of alpha-KG-dependent histone and DNA demethylases by fumarate and succinate that are accumulated in mutations of FH and SDH tumor suppressors, Genes Dev. 26 (12) (2012) 1326–1338. [35] T. Laukka, C.J. Mariani, T. Ihantola, J.Z. Cao, J. Hokkanen, W.G. Kaelin Jr., L. A. Godley, P. Koivunen, Fumarate and succinate regulate expression of hypoxia-inducible genes via TET enzymes, J. Biol. Chem. (2015). [36] E. Letouze, C. Martinelli, C. Loriot, N. Burnichon, N. Abermil, C. Ottolenghi, M. Janin, M. Menara, A.T. Nguyen, P. Benit, A. Buffet, C. Marcaillou, J. Bertherat, L. Amar, P. Rustin, A. De Reynies, A.P. Gimenez-Roqueplo, J. Favier, SDH mutations establish a hypermethylator phenotype in paraganglioma, Cancer Cell 23 (6) (2013) 739–752. [37] J.K. Killian, S.Y. Kim, M. Miettinen, C. Smith, M. Merino, M. Tsokos, M. Quezado, W.I. Smith Jr., M.S. Jahromi, P. Xekouki, E. Szarek, R.L. Walker, J. Lasota, M. Raffeld, B. Klotzle, Z. Wang, L. Jones, Y. Zhu, Y. Wang, J.J. Waterfall, M. J. O’Sullivan, M. Bibikova, K. Pacak, C. Stratakis, K.A. Janeway, J.D. Schiffman, J. B. Fan, L. Helman, P.S. Meltzer, Succinate dehydrogenase mutation underlies global epigenomic divergence in gastrointestinal stromal tumor, Cancer Discov. 3 (6) (2013) 648–657. [38] C. Loriot, M. Domingues, A. Berger, M. Menara, M. Ruel, A. Morin, L.J. CastroVega, E. Letouze, C. Martinelli, A.P. Bemelmans, L. Larue, A.P. Gimenez-Roqueplo, J. Favier, Deciphering the molecular basis of invasiveness in Sdhbdeficient cells, Oncotarget 6 (32) (2015) 32955–32965. [39] J.F. Kerrigan, K.A. Aleck, T.J. Tarby, C.R. Bird, R.A. Heidenreich, Fumaric aciduria: clinical and imaging features, Ann. Neurol. 47 (5) (2000) 583–588. [40] C. Frezza, P.J. Pollard, E. Gottlieb, Inborn and acquired metabolic defects in cancer, J. Mol. Med. 89 (3) (2011) 213–220. [41] L.J. Castro-Vega, A. Buffet, A.A. De Cubas, A. Cascon, M. Menara, E. Khalifa, L. Amar, S. Azriel, I. Bourdeau, O. Chabre, M. Curras-Freixes, V. Franco-Vidal, M. Guillaud-Bataille, C. Simian, A. Morin, R. Leton, A. Gomez-Grana, P.J. Pollard, P. Rustin, M. Robledo, J. Favier, A.P. Gimenez-Roqueplo, Germline mutations in FH confer predisposition to malignant pheochromocytomas and paragangliomas, Hum. Mol. Genet. 23 (9) (2014) 2440–2446. [42] G.R. Clark, M. Sciacovelli, E. Gaude, D.M. Walsh, G. Kirby, M.A. Simpson, R. C. Trembath, J.N. Berg, E.R. Woodward, E. Kinning, P.J. Morrison, C. Frezza, E. R. Maher, Germline FH mutations presenting with pheochromocytoma, J. Clin. Endocrinol. Metab. 99 (10) (2014) E2046–E2050. [43] Y.S. Ha, Y. Chihara, H.Y. Yoon, Y.J. Kim, T.H. Kim, S.H. Woo, S.J. Yun, I.Y. Kim, Y. Hirao, W.J. Kim, Downregulation of fumarate hydratase is related to tumorigenesis in sporadic renal cell cancer, Urol. Int. 90 (2) (2013) 233–239. [44] A. Fieuw, C. Kumps, A. Schramm, F. Pattyn, B. Menten, F. Antonacci, P. Sudmant, J.H. Schulte, N. Van Roy, S. Vergult, P.G. Buckley, A. De Paepe, R. Noguera, R. Versteeg, R. Stallings, A. Eggert, J. Vandesompele, K. De Preter, M. Sciacovelli, C. Frezza / Free Radical Biology and Medicine 100 (2016) 175–181 179 F. Speleman, Identification of a novel recurrent 1q42.2-1qter deletion in high risk MYCN single copy 11q deleted neuroblastomas, Int. J. Cancer 130 (11) (2012) 2599–2606. [45] S. Picaud, K.L. Kavanagh, W.W. Yue, W.H. Lee, S. Muller-Knapp, O. Gileadi, J. Sacchettini, U. Oppermann, Structural basis of fumarate hydratase defi- ciency, J. Inherit. Metab. Dis. 34 (3) (2011) 671–676. [46] C. Frezza, L. Zheng, O. Folger, K.N. Rajagopalan, E.D. MacKenzie, L. Jerby, M. Micaroni, B. Chaneton, J. Adam, A. Hedley, G. Kalna, I.P. Tomlinson, P. J. Pollard, D.G. Watson, R.J. Deberardinis, T. Shlomi, E. Ruppin, E. Gottlieb, Haem oxygenase is synthetically lethal with the tumour suppressor fumarate hydratase, Nature 477 (7363) (2011) 225–228. [47] J. Adam, M. Yang, C. Bauerschmidt, M. Kitagawa, L. O’Flaherty, P. Maheswaran, G. Ozkan, N. Sahgal, D. Baban, K. Kato, K. Saito, K. Iino, K. Igarashi, M. Stratford, C. Pugh, D.A. Tennant, C. Ludwig, B. Davies, P.J. Ratcliffe, M. El-Bahrawy, H. Ashrafian, T. Soga, P.J. Pollard, A role for cytosolic fumarate hydratase in urea cycle metabolism and renal neoplasia, Cell Rep. 3 (5) (2013) 1440–1448. [48] L. Zheng, E.D. Mackenzie, S.A. Karim, A. Hedley, K. Blyth, G. Kalna, D.G. Watson, P. Szlosarek, C. Frezza, E. Gottlieb, Reversed argininosuccinate lyase activity in fumarate hydratase-deficient cancer cells, Cancer Metab. 1 (1) (2013) 12. [49] C.Y. Yoon, Y.J. Shim, E.H. Kim, J.H. Lee, N.H. Won, J.H. Kim, I.S. Park, D.K. Yoon, B. H. Min, Renal cell carcinoma does not express argininosuccinate synthetase and is highly sensitive to arginine deprivation via arginine deiminase, Int. J. Cancer 120 (4) (2007) 897–905. [50] S. Rabinovich, L. Adler, K. Yizhak, A. Sarver, A. Silberman, S. Agron, N. Stettner, Q. Sun, A. Brandis, D. Helbling, S. Korman, S. Itzkovitz, D. Dimmock, I. Ulitsky, S.C. Nagamani, E. Ruppin, A. Erez, Diversion of aspartate in ASS1-deficient tumours fosters de novo pyrimidine synthesis, Nature 527 (7578) (2015) 379–383. [51] M. Yang, T. Soga, P.J. Pollard, J. Adam, The emerging role of fumarate as an oncometabolite, Front. Oncol. 2 (2012) 85. [52] J.S. Isaacs, Y.J. Jung, D.R. Mole, S. Lee, C. Torres-Cabala, Y.L. Chung, M. Merino, J. Trepel, B. Zbar, J. Toro, P.J. Ratcliffe, W.M. Linehan, L. Neckers, HIF overexpression correlates with biallelic loss of fumarate hydratase in renal cancer: novel role of fumarate in regulation of HIF stability, Cancer Cell 8 (2) (2005) 143–153. [53] K. Shanmugasundaram, B. Nayak, E.H. Shim, C.B. Livi, K. Block, S. Sudarshan, The oncometabolite fumarate promotes pseudohypoxia through noncanonical activation of NF-kappaB signaling, J. Biol. Chem. 289 (35) (2014) 24691–24699. [54] J. Adam, E. Hatipoglu, L. O’Flaherty, N. Ternette, N. Sahgal, H. Lockstone, D. Baban, E. Nye, G.W. Stamp, K. Wolhuter, M. Stevens, R. Fischer, P. Carmeliet, P.H. Maxwell, C.W. Pugh, N. Frizzell, T. Soga, B.M. Kessler, M. El-Bahrawy, P. J. Ratcliffe, P.J. Pollard, Renal cyst formation in Fh1-deficient mice is independent of the Hif/Phd pathway: roles for fumarate in KEAP1 succination and Nrf2 signaling, Cancer Cell 20 (4) (2011) 524–537. [55] C. Sourbier, C.J. Ricketts, S. Matsumoto, D.R. Crooks, P.J. Liao, P.Z. Mannes, Y. Yang, M.H. Wei, G. Srivastava, S. Ghosh, V. Chen, C.D. Vocke, M. Merino, R. Srinivasan, M.C. Krishna, J.B. Mitchell, A.M. Pendergast, T.A. Rouault, L. Neckers, W.M. Linehan, Targeting ABL1-mediated oxidative stress adaptation in fumarate hydratase-deficient cancer, Cancer Cell 26 (6) (2014) 840–850. [56] W.M. Linehan, P.T. Spellman, C.J. Ricketts, C.J. Creighton, S.S. Fei, C. Davis, D. A. Wheeler, B.A. Murray, L. Schmidt, C.D. Vocke, M. Peto, A.A. Al Mamun, E. Shinbrot, A. Sethi, S. Brooks, W.K. Rathmell, A.N. Brooks, K.A. Hoadley, A. G. Robertson, D. Brooks, R. Bowlby, S. Sadeghi, H. Shen, D.J. Weisenberger, M. Bootwalla, S.B. Baylin, P.W. Laird, A.D. Cherniack, G. Saksena, S. Haake, J. Li, H. Liang, Y. Lu, G.B. Mills, R. Akbani, M.D. Leiserson, B.J. Raphael, P. Anur, D. Bottaro, L. Albiges, N. Barnabas, T.K. Choueiri, B. Czerniak, A.K. Godwin, A. A. Hakimi, T.H. Ho, J. Hsieh, M. Ittmann, W.Y. Kim, B. Krishnan, M.J. Merino, K. R. Mills Shaw, V.E. Reuter, E. Reznik, C.S. Shelley, B. Shuch, S. Signoretti, R. Srinivasan, P. Tamboli, G. Thomas, S. Tickoo, K. Burnett, D. Crain, J. Gardner, K. Lau, D. Mallery, S. Morris, J.D. Paulauskis, R.J. Penny, C. Shelton, W. T. Shelton, M. Sherman, E. Thompson, P. Yena, M.T. Avedon, J. Bowen, J. M. Gastier-Foster, M. Gerken, K.M. Leraas, T.M. Lichtenberg, N.C. Ramirez, T. Santos, L. Wise, E. Zmuda, J.A. Demchok, I. Felau, C.M. Hutter, M. Sheth, H. J. Sofia, R. Tarnuzzer, Z. Wang, L. Yang, J.C. Zenklusen, J. Zhang, B. Ayala, J. Baboud, S. Chudamani, J. Liu, L. Lolla, R. Naresh, T. Pihl, Q. Sun, Y. Wan, Y. Wu, A. Ally, M. Balasundaram, S. Balu, R. Beroukhim, T. Bodenheimer, C. Buhay, Y. S. Butterfield, R. Carlsen, S.L. Carter, H. Chao, E. Chuah, A. Clarke, K. R. Covington, M. Dahdouli, N. Dewal, N. Dhalla, H.V. Doddapaneni, J. A. Drummond, S.B. Gabriel, R.A. Gibbs, R. Guin, W. Hale, A. Hawes, D.N. Hayes, R.A. Holt, A.P. Hoyle, S.R. Jefferys, S.J. Jones, C.D. Jones, D. Kalra, C. Kovar, L. Lewis, Y. Ma, M.A. Marra, M. Mayo, S. Meng, M. Meyerson, P.A. Mieczkowski, R.A. Moore, D. Morton, L.E. Mose, A.J. Mungall, D. Muzny, J.S. Parker, C. M. Perou, J. Roach, J.E. Schein, S.E. Schumacher, Y. Shi, J.V. Simons, P. Sipahimalani, T. Skelly, M.G. Soloway, C. Sougnez, A. Tam, D. Tan, N. Thiessen, U. Veluvolu, M. Wang, M.D. Wilkerson, T. Wong, J. Wu, L. Xi, J. Zhou, J. Bedford, F. Chen, Y. Fu, M. Gerstein, D. Haussler, K. Kasaian, P. Lai, S. Ling, A. Radenbaugh, D. Van Den Berg, J.N. Weinstein, J. Zhu, M. Albert, I. Alexopoulou, J.J. Andersen, J.T. Auman, J. Bartlett, S. Bastacky, J. Bergsten, M. L. Blute, L. Boice, R.J. Bollag, J. Boyd, E. Castle, Y.B. Chen, J.C. Cheville, E. Curley, B. Davies, A. DeVolk, R. Dhir, L. Dike, J. Eckman, J. Engel, J. Harr, R. Hrebinko, M. Huang, L. Huelsenbeck-Dill, M. Iacocca, B. Jacobs, M. Lobis, J.K. Maranchie, S. McMeekin, J. Myers, J. Nelson, J. Parfitt, A. Parwani, N. Petrelli, B. Rabeno, S. Roy, A.L. Salner, J. Slaton, M. Stanton, R.H. Thompson, L. Thorne, K. Tucker, P. M. Weinberger, C. Winemiller, L.A. Zach, R. Zuna, Comprehensive molecular characterization of papillary renal-cell carcinoma, N. Engl. J. Med. 374 (2) (2016) 135–145. [57] N.L. Alderson, Y. Wang, M. Blatnik, N. Frizzell, M.D. Walla, T.J. Lyons, N. Alt, J. A. Carson, R. Nagai, S.R. Thorpe, J.W. Baynes, S-(2-Succinyl)cysteine: a novel chemical modification of tissue proteins by a Krebs cycle intermediate, Arch. Biochem. Biophys. 450 (1) (2006) 1–8. [58] C. Bardella, M. El-Bahrawy, N. Frizzell, J. Adam, N. Ternette, E. Hatipoglu, K. Howarth, L. O’Flaherty, I. Roberts, G. Turner, J. Taylor, K. Giaslakiotis, V. M. Macaulay, A.L. Harris, A. Chandra, H.J. Lehtonen, V. Launonen, L.A. Aaltonen, C.W. Pugh, R. Mihai, D. Trudgian, B. Kessler, J.W. Baynes, P.J. Ratcliffe, I. P. Tomlinson, P.J. Pollard, Aberrant succination of proteins in fumarate hydratase-deficient mice and HLRCC patients is a robust biomarker of mutation status, J. Pathol. 225 (1) (2011) 4–11. [59] N. Frizzell, S.A. Thomas, J.A. Carson, J.W. Baynes, Mitochondrial stress causes increased succination of proteins in adipocytes in response to glucotoxicity, Biochem. J. 445 (2) (2012) 247–254. [60] M. Yang, N. Ternette, H. Su, R. Dabiri, B.M. Kessler, J. Adam, B.T. Teh, P.J. Pollard, The Succinated Proteome of FH-Mutant Tumours, Metabolites 4 (3) (2014) 640–654. [61] N. Ternette, M. Yang, M. Laroyia, M. Kitagawa, L. O’Flaherty, K. Wolhulter, K. Igarashi, K. Saito, K. Kato, R. Fischer, A. Berquand, B.M. Kessler, T. Lappin, N. Frizzell, T. Soga, J. Adam, P.J. Pollard, Inhibition of mitochondrial aconitase by succination in fumarate hydratase deficiency, Cell Rep. 3 (3) (2013) 689–700. [62] A. Ooi, J.C. Wong, D. Petillo, D. Roossien, V. Perrier-Trudova, D. Whitten, B. W. Min, M.H. Tan, Z. Zhang, X.J. Yang, M. Zhou, B. Gardie, V. Molinie, S. Richard, P.H. Tan, B.T. Teh, K.A. Furge, An antioxidant response phenotype shared between hereditary and sporadic type 2 papillary renal cell carcinoma, Cancer Cell 20 (4) (2011) 511–523. [63] G.M. DeNicola, F.A. Karreth, T.J. Humpton, A. Gopinathan, C. Wei, K. Frese, D. Mangal, K.H. Yu, C.J. Yeo, E.S. Calhoun, F. Scrimieri, J.M. Winter, R.H. Hruban, C. Iacobuzio-Donahue, S.E. Kern, I.A. Blair, D.A. Tuveson, Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis, Nature 475 (7354) (2011) 106–109. [64] L.B. Sullivan, E. Martinez-Garcia, H. Nguyen, A.R. Mullen, E. Dufour, S. Sudarshan, J.D. Licht, R.J. Deberardinis, N.S. Chandel, The proto-oncometabolite fumarate binds glutathione to amplify ROS-dependent signaling, Mol. Cell 51 (2) (2013) 236–248. [65] L. Zheng, S. Cardaci, L. Jerby, E.D. MacKenzie, M. Sciacovelli, T.I. Johnson, E. Gaude, A. King, J.D. Leach, R. Edrada-Ebel, A. Hedley, N.A. Morrice, G. Kalna, K. Blyth, E. Ruppin, C. Frezza, E. Gottlieb, Fumarate induces redox-dependent senescence by modifying glutathione metabolism, Nat. Commun. 6 (2015) 6001. [66] L. Jin, D. Li, G.N. Alesi, J. Fan, H.B. Kang, Z. Lu, T.J. Boggon, P. Jin, H. Yi, E. R. Wright, D. Duong, N.T. Seyfried, R. Egnatchik, R.J. DeBerardinis, K. R. Magliocca, C. He, M.L. Arellano, H.J. Khoury, D.M. Shin, F.R. Khuri, S. Kang, Glutamate dehydrogenase 1 signals through antioxidant glutathione peroxidase 1 to regulate redox homeostasis and tumor growth, Cancer Cell 27 (2) (2015) 257–270. [67] A.R. Mullen, R.J. DeBerardinis, Genetically-defined metabolic reprogramming in cancer, Trends Endocrinol. Metab. 23 (11) (2012) 552–559. [68] L. Dang, D.W. White, S. Gross, B.D. Bennett, M.A. Bittinger, E.M. Driggers, V. R. Fantin, H.G. Jang, S. Jin, M.C. Keenan, K.M. Marks, R.M. Prins, P.S. Ward, K. E. Yen, L.M. Liau, J.D. Rabinowitz, L.C. Cantley, C.B. Thompson, M.G. Vander Heiden, S.M. Su, Cancer-associated IDH1 mutations produce 2-hydroxyglutarate, Nature 462 (7274) (2009) 739–744. [69] P.S. Ward, J. Patel, D.R. Wise, O. Abdel-Wahab, B.D. Bennett, H.A. Coller, J. R. Cross, V.R. Fantin, C.V. Hedvat, A.E. Perl, J.D. Rabinowitz, M. Carroll, S.M. Su, K.A. Sharp, R.L. Levine, C.B. Thompson, The common feature of leukemia-associated IDH1 and IDH2 mutations is a neomorphic enzyme activity converting alpha-ketoglutarate to 2-hydroxyglutarate, Cancer Cell 17 (3) (2010) 225–234. [70] C.L. Linster, E. Van Schaftingen, A.D. Hanson, Metabolite damage and its repair or pre-emption, Nat. Chem. Biol. 9 (2) (2013) 72–80. [71] M. Aghili, F. Zahedi, E. Rafiee, Hydroxyglutaric aciduria and malignant brain tumor: a case report and literature review, J. Neurooncol. 91 (2) (2009) 233–236. [72] M. Kranendijk, E.A. Struys, G.S. Salomons, M.S. Van der Knaap, C. Jakobs, Progress in understanding 2-hydroxyglutaric acidurias, J. Inherit. Metab. Dis. 35 (4) (2012) 571–587. [73] I. Moroni, M. Bugiani, L. D’Incerti, C. Maccagnano, M. Rimoldi, L. Bissola, B. Pollo, G. Finocchiaro, G. Uziel, L-2-hydroxyglutaric aciduria and brain malignant tumors: a predisposing condition? Neurology 62 (10) (2004) 1882–1884. [74] E.H. Shim, C.B. Livi, D. Rakheja, J. Tan, D. Benson, V. Parekh, E.Y. Kho, A. P. Ghosh, R. Kirkman, S. Velu, S. Dutta, B. Chenna, S.L. Rea, R.J. Mishur, Q. Li, T. L. Johnson-Pais, L. Guo, S. Bae, S. Wei, K. Block, S. Sudarshan, L-2-Hydroxyglutarate: an epigenetic modifier and putative oncometabolite in renal cancer, Cancer Discov. 4 (11) (2014) 1290–1298. [75] P. Koivunen, S. Lee, C.G. Duncan, G. Lopez, G. Lu, S. Ramkissoon, J.A. Losman, P. Joensuu, U. Bergmann, S. Gross, J. Travins, S. Weiss, R. Looper, K.L. Ligon, R. G. Verhaak, H. Yan, W.G. Kaelin Jr., Transformation by the (R)-enantiomer of 2-hydroxyglutarate linked to EGLN activation, Nature 483 (7390) (2012) 484–488. [76] R. Chowdhury, K.K. Yeoh, Y.M. Tian, L. Hillringhaus, E.A. Bagg, N.R. Rose, I. K. Leung, X.S. Li, E.C. Woon, M. Yang, M.A. McDonough, O.N. King, I.J. Clifton, R. J. Klose, T.D. Claridge, P.J. Ratcliffe, C.J. Schofield, A. Kawamura, The 180 M. Sciacovelli, C. Frezza / Free Radical Biology and Medicine 100 (2016) 175–181 oncometabolite 2-hydroxyglutarate inhibits histone lysine demethylases, EMBO Rep. 12 (5) (2011) 463–469. [77] W. Xu, H. Yang, Y. Liu, Y. Yang, P. Wang, S.H. Kim, S. Ito, C. Yang, M.T. Xiao, L. X. Liu, W.Q. Jiang, J. Liu, J.Y. Zhang, B. Wang, S. Frye, Y. Zhang, Y.H. Xu, Q.Y. Lei, K.L. Guan, S.M. Zhao, Y. Xiong, Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of alpha-ketoglutarate-dependent dioxygenases, Cancer Cell 19 (1) (2011) 17–30. [78] S. Turcan, D. Rohle, A. Goenka, L.A. Walsh, F. Fang, E. Yilmaz, C. Campos, A. W. Fabius, C. Lu, P.S. Ward, C.B. Thompson, A. Kaufman, O. Guryanova, R. Levine, A. Heguy, A. Viale, L.G. Morris, J.T. Huse, I.K. Mellinghoff, T.A. Chan, IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype, Nature 483 (7390) (2012) 479–483. [79] M.E. Figueroa, O. Abdel-Wahab, C. Lu, P.S. Ward, J. Patel, A. Shih, Y. Li, N. Bhagwat, A. Vasanthakumar, H.F. Fernandez, M.S. Tallman, Z. Sun, K. Wolniak, J.K. Peeters, W. Liu, S.E. Choe, V.R. Fantin, E. Paietta, B. Lowenberg, J.D. Licht, L.A. Godley, R. Delwel, P.J. Valk, C.B. Thompson, R.L. Levine, A. Melnick, Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation, Cancer Cell 18 (6) (2010) 553–567. [80] C. Lu, P.S. Ward, G.S. Kapoor, D. Rohle, S. Turcan, O. Abdel-Wahab, C. R. Edwards, R. Khanin, M.E. Figueroa, A. Melnick, K.E. Wellen, D.M. O’Rourke, S. L. Berger, T.A. Chan, R.L. Levine, I.K. Mellinghoff, C.B. Thompson, IDH mutation impairs histone demethylation and results in a block to cell differentiation, Nature 483 (7390) (2012) 474–478. [81] W.A. Flavahan, Y. Drier, B.B. Liau, S.M. Gillespie, A.S. Venteicher, A.O. StemmerRachamimov, M.L. Suva, B.E. Bernstein, Insulator dysfunction and oncogene activation in IDH mutant gliomas, Nature 529 (7584) (2016) 110–114. [82] X. Fu, R.M. Chin, L. Vergnes, H. Hwang, G. Deng, Y. Xing, M.Y. Pai, S. Li, L. Ta, F. Fazlollahi, C. Chen, R.M. Prins, M.A. Teitell, D.A. Nathanson, A. Lai, K.F. Faull, M. Jiang, S.G. Clarke, T.F. Cloughesy, T.G. Graeber, D. Braas, H.R. Christofk, M. E. Jung, K. Reue, J. Huang, 2-Hydroxyglutarate Inhibits ATP Synthase and mTOR Signaling, Cell Metab. 22 (3) (2015) 508–515. [83] J.A. Losman, R.E. Looper, P. Koivunen, S. Lee, R.K. Schneider, C. McMahon, G. S. Cowley, D.E. Root, B.L. Ebert, W.G. Kaelin Jr., (R)-2-hydroxyglutarate is suf- ficient to promote leukemogenesis and its effects are reversible, Science 339 (6127) (2013) 1621–1625. [84] A. Terunuma, N. Putluri, P. Mishra, E.A. Mathe, T.H. Dorsey, M. Yi, T.A. Wallace, H.J. Issaq, M. Zhou, J.K. Killian, H.S. Stevenson, E.D. Karoly, K. Chan, S. Samanta, D. Prieto, T.Y. Hsu, S.J. Kurley, V. Putluri, R. Sonavane, D.C. Edelman, J. Wulff, A. M. Starks, Y. Yang, R.A. Kittles, H.G. Yfantis, D.H. Lee, O.B. Ioffe, R. Schiff, R. M. Stephens, P.S. Meltzer, T.D. Veenstra, T.F. Westbrook, A. Sreekumar, S. Ambs, MYC-driven accumulation of 2-hydroxyglutarate is associated with breast cancer prognosis, J. Clin. Invest. 124 (1) (2014) 398–412. [85] J. Fan, X. Teng, L. Liu, K.R. Mattaini, R.E. Looper, M.G. Vander Heiden, J. D. Rabinowitz, Human phosphoglycerate dehydrogenase produces the oncometabolite d-2-hydroxyglutarate, ACS Chem. Biol. 10 (2) (2015) 510–516. [86] F. Li, X. He, D. Ye, Y. Lin, H. Yu, C. Yao, L. Huang, J. Zhang, F. Wang, S. Xu, X. Wu, L. Liu, C. Yang, J. Shi, J. Liu, Y. Qu, F. Guo, J. Zhao, W. Xu, S. Zhao, NADP(+)-IDH Mutations Promote Hypersuccinylation that Impairs Mitochondria Respiration and Induces Apoptosis Resistance, Mol. Cell. 60 (4) (2015) 661–675. [87] D.R. Wise, P.S. Ward, J.E. Shay, J.R. Cross, J.J. Gruber, U.M. Sachdeva, J.M. Platt, R. G. DeMatteo, M.C. Simon, C.B. Thompson, Hypoxia promotes isocitrate dehydrogenase-dependent carboxylation of alpha-ketoglutarate to citrate to support cell growth and viability, Proc. Natl. Acad. Sci. USA 108 (49) (2011) 19611–19616. [88] A.M. Intlekofer, R.G. Dematteo, S. Venneti, L.W. Finley, C. Lu, A.R. Judkins, A. S. Rustenburg, P.B. Grinaway, J.D. Chodera, J.R. Cross, C.B. Thompson, Hypoxia induces production of L-2-hydroxyglutarate, Cell Metab. 22 (2) (2015) 304–311. [89] W.M. Oldham, C.B. Clish, Y. Yang, J. Loscalzo, Hypoxia-mediated increases in L-2-hydroxyglutarate coordinate the metabolic response to reductive stress, Cell Metab. 22 (2) (2015) 291–303. [90] W.R. Wilson, M.P. Hay, Targeting hypoxia in cancer therapy, Nat. Rev. Cancer 11 (6) (2011) 393–410. [91] E.T. Chouchani, V.R. Pell, E. Gaude, D. Aksentijevic, S.Y. Sundier, E.L. Robb, A. Logan, S.M. Nadtochiy, E.N. Ord, A.C. Smith, F. Eyassu, R. Shirley, C.H. Hu, A. J. Dare, A.M. James, S. Rogatti, R.C. Hartley, S. Eaton, A.S. Costa, P.S. Brookes, S. M. Davidson, M.R. Duchen, K. Saeb-Parsy, M.J. Shattock, A.J. Robinson, L. M. Work, C. Frezza, T. Krieg, M.P. Murphy, Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS, Nature 515 (7527) (2014) 431–435. [92] D. Rodenhizer, E. Gaude, D. Cojocari, R. Mahadevan, C. Frezza, B.G. Wouters, A. P. McGuigan, A three-dimensional engineered tumour for spatial snapshot analysis of cell metabolism and phenotype in hypoxic gradients, Nat. Mater. 15 (2) (2016) 227–234. [93] M. Blatnik, S.R. Thorpe, J.W. Baynes, Succination of proteins by fumarate: mechanism of inactivation of glyceraldehyde-3-phosphate dehydrogenase in diabetes, Ann. N. Y. Acad. Sci. 1126 (2008) 272–275. [94] T. Nishikawa, D. Edelstein, X.L. Du, S. Yamagishi, T. Matsumura, Y. Kaneda, M. A. Yorek, D. Beebe, P.J. Oates, H.P. Hammes, I. Giardino, M. Brownlee, Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage, Nature 404 (6779) (2000) 787–790. [95] S.A. Thomas, K.B. Storey, J.W. Baynes, N. Frizzell, Tissue distribution of S-(2- succino)cysteine (2SC), a biomarker of mitochondrial stress in obesity and diabetes, Obesity 20 (2) (2012) 263–269. M. Sciacovelli, C. Frezza / Free Radical Biology and Medicine 100 (2016) 175–181 181