Open access peer-reviewed chapter

IDH-Mutant Gliomas

Written By

Kensuke Tateishi and Tetsuya Yamamoto

Submitted: 08 October 2018 Reviewed: 18 January 2019 Published: 23 February 2019

DOI: 10.5772/intechopen.84543

From the Edited Volume

Brain and Spinal Tumors - Primary and Secondary

Edited by Lee Roy Morgan and Feyzi Birol Sarica

Chapter metrics overview

2,238 Chapter Downloads

View Full Metrics

Abstract

Isocitrate dehydrogenase (IDH) mutation is one of the most critical genomic alterations in lower grade and secondary glioblastoma patient. More than 90% of IDH mutation is located at codon R132 of IDH1 gene. IDH mutation produces oncometabolite “2-hydroxyglutarate” and induces epigenetic alteration, such as DNA global methylation and histone methylation. As a result, IDH mutation promotes early gliomagenesis. Since IDH mutation is the earliest genomic event and almost always retained during tumor progression, IDH mutation is expected as novel therapeutic target. Herein, we review the clinical characteristics of IDH-mutant gliomas, biological role of IDH mutation for gliomagenesis, and current and future therapeutic approach for IDH mutant tumors.

Keywords

  • IDH mutation
  • glioma
  • 2-hydroxyglutarate
  • tumor biology
  • cancer metabolism
  • target therapy

1. Introduction

The WHO 2016 classification integrates molecular and histological features in the diagnosis of gliomas. Among numerous genomic alterations, the isocitrate dehydrogenase (IDH) mutation is one of the most important genetic alterations found in this kind of tumor. As IDH mutation is a ubiquitous mutation in lower grade gliomas, the development of molecular target therapies against IDH mutations is expected. Here, we review IDH-mutant gliomas, focusing on their role in tumorigenesis and as novel therapeutic targets.

Advertisement

2. Discovery of IDH mutations in cancers

The presence of an isocitrate dehydrogenase (IDH) mutation was first discovered in colorectal cancers [1]. Parsons et al. [2] found mutations of the IDH1 (2q.33) in 12% of the glioblastomas (GBMs). Other large scale studies validated that IDH1 and IDH2 (IDH) mutations were found in the majority of secondary GBM and lower grade (WHO grade II and III) gliomas, whereas these were rarely found in adult primary and pediatric GBMs [2, 3, 4]. Almost all of the IDH1 mutations occur at codon 132, >90% of them exhibit a c.395G>A (R132H) substitution, followed by R132C [3, 5, 6]. Although the frequency was low, IDH2 mutations were also identified at codon 172 in gliomas [4, 7].

Besides, IDH mutation was found in hematopoietic cancers, including acute myeloid leukemia (AML; 10–15%, IDH2) [8, 9], angioimmunoblastic T-cell lymphoma (AITL, 20%) [10], chondrosarcoma (~50%) [11, 12, 13], intrahepatic cholangiocarcinoma (15–20%, IDH1) [13], and at lower frequency in other hematopoietic and solid cancers, such as B-acute lymphoblastic leukemia (B-ALL), esophageal cancer, colorectal cancer, melanoma, prostate carcinoma, and breast adenocarcinoma [1, 4, 14, 15, 16].

Advertisement

3. Tumorigenesis of IDH-mutant gliomas

3.1 Genomic characteristics of IDH-mutant glioma

The discovery of IDH mutations allowed the distinction of primary GBM, which is genetically characterized by TERT promoter mutation, gene alteration of epidermal growth factor receptor (EGFR), phosphatase and tensin homolog (PTEN) mutation or deletion, trisomy 7, monosomy 10, and cyclin-dependent kinase inhibitor 2A (CDKN2A) homozygous deletion, from secondary GBM (GBM, IDH-mutant) [3, 5, 17, 18].

In astrocytic tumors, most of the tumor cells have co-mutations in IDH1, TP53, and ATRX. Moreover, WHO 2016 [19] defined the presence of IDH mutation and co-deletion of chromosome1p and 19q as necessary for the diagnosis of oligodendroglial tumors. Also, in oligodendroglial tumors, TERT promoter mutation is almost always present (>95%), while CIC and FUBP1 are commonly (>40%) observed. These genetic abnormalities for astrocytic and oligodendroglial tumors are mutually exclusive [20, 21, 22, 23, 24]. Importantly, the IDH mutation is the earliest genetic alteration observed; it is commonly retained during tumor progression [25, 26, 27, 28], but in a subset of mutants, IDH1 was either deleted or amplified at tumor recurrence [29], indicating the critical role of IDH mutation for tumorigenesis. It has also been shown that IDH mutations do not select or create ATRX or TERT promoter mutations [30].

3.2 Developmental hierarchy in IDH-mutant gliomas

Two recent large scale single cell RNA-sequencing studies revealed a developmental hierarchy in IDH1-mutant gliomas [31, 32]. Accordingly, IDH1-mutant astrocytoma and oligodendroglioma shared a similar developmental hierarchy, consisting of three subpopulations of malignant cells: nonproliferative astrocytic and oligodendrocytic cells, proliferative, and undifferentiated neural stem/progenitor cells. In contrast, tumor micro environment (TME) was different in the abundance microglia/macrophage cells between astrocytic and oligodendroglial tumors. TME also differs between astrocytic tumors of different grades. Though TME and genomic alterations are distinctive, evidence indicates the existence of common progenitor cells in IDH1-mutant gliomas. In higher grade tumors, undifferentiated glioma stem/progenitor cells were increased [32]. In addition, almost all proliferating cancer cells were composed of subpopulations of undifferentiated cells (stem-like) in oligodendroglioma [31], suggesting a significant role for undifferentiated cells in cell proliferation and malignant progression.

3.3 IDH-mutant xenograft model

Although IDH1 mutation induced proliferation in vitro [33], IDH1 mutation did not promote xenograft formation [34, 35, 36]. Intriguingly, Bardella et al. [37] demonstrated that IDH1R132H overexpression in the murine subventricular zone induced the formation of early gliomagenesis, where stem and transit amplifying/progenitor cell populations were expanded, indicating the pivotal role of IDH1 mutation in glioma formation. Moreover, Wakimoto et al. demonstrated that “tertiary mutations,” such as PIK3CA mutation, PDGFRA amplification, and MYC amplification, promote IDH1-mutant glioma formation in xenograft models. Importantly, tumor harboring tertiary mutations were associated with unfavorable prognosis in IDH1-mutant glioma patients [38]. Recently, large genomic analyses demonstrated that malignant progression in IDH1-mutant glioma is associated with the PI3K pathway and MYC activation [39, 40]. Thus, IDH mutation induces gliomagenesis, whereas tertiary mutations are critical to promote tumor progression in lower grade gliomas (Figure 1).

Figure 1.

Genomic alteration and tumor microenvironment in IDH-mutant astrocytic and oligodendroglial tumors.

Advertisement

4. The 2016 WHO classification

The 2016 World Health Organization (WHO) Classification of Tumors of the Central Nervous System (CNS) integrated phenotypic and genotypic parameters for CNS tumor classification. According to this classification, all diffusely infiltrating gliomas are grouped as diffuse astrocytic and oligodendroglial tumors. These tumors were histologically and genetically classified based on the presence of IDH mutation, co-deletion of chromosome1p and 19q, or ATRX and TP53 mutations. Accordingly, gliomas are classified as follows: (1) diffuse astrocytoma (WHO grade II) or anaplastic astrocytoma (AA, WHO grade III): IDH-mutant, -wildtype, or not otherwise specified (NOS); (2) oligodendroglioma (WHO grade II) or anaplastic oligodendroglioma (WHO grade III): IDH-mutant and 1p/19q-codeleted or NOS; (3) oligoastrocytoma (grade II) and anaplastic oligoastrocytoma (WHO grade III): NOS; (4) GBM (WHO grade IV): IDH-mutant, -wildtype, or NOS; and (5) diffuse midline glioma (WHO grade IV): H3K27M-mutant.

IDH-wildtype GBM (about 90% of cases) is known as primary GBM, while IDH-mutant GBM (about 10% of cases) corresponds to secondary GBM [19].

Advertisement

5. Epidemiology of IDH-mutant gliomas

5.1 Age distribution of IDH-mutant gliomas

According to some statistical analyses, the IDH-mutant GBM or anaplastic astrocytoma patients were more than 20 years younger than those with IDH-wildtype GBM [4]. In contrast, IDH-mutant GBM patients were only 4 years older than those with IDH1-mutant grade II and III astrocytoma [41]. This indicates that IDH-mutant glioma arises earlier than IDH-wildtype glioma (mostly GBM).

5.2 Prognosis of IDH-mutant gliomas

Parsons et al. [2] initially demonstrated that IDH1-mutant GBM patients survived about threefold longer than those with IDH1-wildtype GBM. Other groups verified that IDH1 mutation is a favorable prognostic biomarker in gliomas [4, 42, 43]. In addition to GBM, large amounts of clinical studies indicated that the IDH mutation was an independent prognostic factor in grade II and III gliomas [4, 28, 43, 44, 45, 46, 47]. Notably, the prognosis of IDH1-mutant GBM is better than of IDH1-wildtype AA [48]. Also, a prospective randomized study (NOA-04) revealed that IDH1 mutation, hypermethylation of the O6-methylguanine DNA-methyltransferase (MGMT) promoter, age, extent of resection, and oligodendroglial histology are independent prognostic factors in anaplastic gliomas [44]. Among them, the impact of IDH1 mutation conferred a stronger favorable prognosis than 1p/19q co-deletion, MGMT promoter methylation, and histology [44]. Collectively, IDH1 mutation is a convincing prognostic factor in gliomas, irrespective of tumor grade and histology.

5.3 Prognostic classification for gliomas

Suzuki et al. [28] distinguished lower grade gliomas on the basis of the presence of IDH1 mutation, TP53 mutation, and 1p/19q co-deletion. Accordingly, tumors were classified into three groups: type I (IDH1-mutant with 1p/19q co-deletion; favorable prognostic group), type II (IDH1-mutant with TP53 mutation; intermediate prognostic group), and type III (IDH1-wildtype; poor prognostic group). Eckel-Passow et al. [47] classified gliomas into five groups based on the mutation status of IDH1 and TERT promoter and on 1p/19q co-deletion. This group also demonstrated that TERT promoter mutations and ATRX alterations provided additional information for a tailored prognostic classification [49]. Besides, Arita et al. [50] proposed a classification of grade II–IV gliomas based on the mutations in IDH and the hotspot in TERT promoter.

Among IDH-mutant astrocytic tumors, CDKN2A/B homozygous deletion was demonstrated to be an unfavorable prognostic molecular marker [51]. Similarly, another group demonstrated that PIK3R1 mutation and altered retinoblastoma pathway genes, including RB1 and CDKN2A, were independent predictors of poor survival in astrocytic tumors. In oligodendrogliomas, NOTCH pathway inactivation and PI3K pathway activation were associated with poor prognosis [52, 53]. Collectively, these molecular markers could predict prognosis in glioma patients.

Advertisement

6. The mechanism of tumorigenesis in IDH1-mutant gliomas

6.1 IDH mutation drives production of oncometabolite D-2-hydroxyglutarate

In humans, IDH is composed of three types of isozymes (IDH1, IDH2, and IDH3). IDH1 is located in the cytoplasm and peroxisomes, whereas IDH2 and IDH3 are localized in the mitochondria and are involved in the TCA cycle. IDH1 and IDH2 are NADP+ dependent, whereas IDH3 is NAD+ dependent. IDH converts isocitrate into α-ketoglutarate (α-KG). No mutation in IDH3 has been detected in human cancers. If IDH is mutated, it blocks normal enzymatic activity and instead produces D-2-hydroxyglutarate (2-HG) from α-KG in an NADPH dependent manner, irrespective of the substituted amino acid [54, 55, 56]. Compared with IDH-wildtype cells, the 2-HG level in IDH-mutant cells is 50–100-fold higher [54, 57]. IDH mutations are almost always heterozygous, and both mutant and wildtype IDH1 alleles are required for 2-HG production in glioma cells [58].

6.2 IDH-mutation induced epigenetic alterations

6.2.1 IDH-mutation inducible DNA hypermethylator phenotype

Since the structure of 2-HG is similar to that of α-KG, 2-HG inhibits a variety of α-KG-dependent dioxygenases [59, 60]. Among them, 10–11 translocation-2 (TET2) induces global demethylation of DNA by catalyzing the conversion of 5-methylcytosine (5-mC) to 5-hydroxymethylcytosine (5-hmC). Forced mutant IDH1 caused increased 5mC concentrations, instead of decreased 5hmC [37, 61]. IDH mutation also promotes methylation of DNA by TET2 inhibition, resulting in a glioma CpG island methylator phenotype (G-CIMP), a specific DNA methylation pattern in IDH-mutant tumor cells [61, 62, 63]. Indeed, forced overexpression of mutant IDH (IDH1R132H and IDH2R172K) produced high concentrations of 2-HG and increased global 5-mC levels [61]. Similarly, TET2 mutations, which are mutually exclusive to IDH mutations, induce a global hypermethylation signature [61]. Turcan et al. [64] demonstrated that a G-CIMP-like phenotype and G-CIMP positive proneural glioblastomas were formed after the introduction of an IDH1 mutation into normal human astrocytes (NHA). These data indicate that mutant IDH induced TET2 suppression, followed by G-CIMP, in cancer cells. Consistent with IDH-mutant glioma patients, glioma patients with G-CIMP are younger at diagnosis and survive longer than those without G-CIMP [62]. Intriguingly, about 10% of G-CIMP tumors were relapsed as G-CIMP low tumors with poor clinical outcome [65].

The Cancer Genome Atlas (TCGA) performed comprehensive transcriptome analysis. Accordingly, GBM was classified into four groups (classic, mesenchymal, proneural, and neural groups). Aberrations and gene expression of EGFR and NF1 define the classical and mesenchymal subtypes, whereas tumors with an IDH1 mutation were classified within the proneural group. The proneural group is also accompanied by a PDGFRA gene abnormality and the G-CIMP feature [66]. DNA methylation induced by the IDH1 mutation caused hypermethylation at cohesion and CCCTC-binding factor (CTCF) binding sites and compromised the binding of the insulator protein. As a result, loss of CTCF at a domain permits a constitutive enhancer to interact aberrantly with the receptor tyrosine kinase gene PDGFRA [67].

6.2.2 IDH mutation promotes global histone methylation

IDH mutation is also known to increase histone methylation. Lysine methylation of histone tails modifies chromatin structure and regulates gene expression. By competition with α-KG, 2-HG inhibits histone demethylases including members of the Jumonji transcription factor family (JMJD2A, JMJD2C/KDM4C, and JHDM1A/FBXL11), resulting in histone hypermethylation [68]. Indeed, hypermethylation in H3K4me1, H3K4me3, H3K9me2, H3K27me2, H3K79me2, H3K27me3, H3K9me3, and H3K36me3 was observed in cells with exogenous 2-HG or mutant IDH1 induction [60, 63, 64, 69]. Sasaki et al. [63] also demonstrated that IDH1R132H knock in mice showed significantly increased early hematopoietic progenitors, histone hypermethylation, and DNA methylation. Interestingly, the elevation of H3K9me3 levels was observed earlier than the DNA methylation change in NHA upon IDH1R132H induction [69], suggesting that histone methylation may be an early event in IDH1-mutant cancers. The hypermethylation of histones blocks cell differentiation in cancer cells [60, 63, 64, 69]. Using a histone demethylating agent or a specific mutant IDH1 inhibitor, suppressed cell differentiation can be restored [70, 71]. Besides, 2-HG impairs collagen maturation, which leads to basement membrane aberrations that play a part in glioma progression [72]. Taken together, these data show that DNA hypermethylation and histone methylation promote tumorigenesis through a wide range of gene function changes (Figure 2).

Figure 2.

Biological role of IDH mutation to induce gliomagenesis.

6.3 IDH mutation inducible metabolic alterations

In addition to the epigenetic changes, IDH1 mutation is known to alter hypoxia inducible factor 1α (HIF-1α) activity. Under oxidative conditions, α-KG-dependent prolyl hydroxylases (PHDs), which form the Egl nine homolog (EglN) families, induce HIF-1α hydroxylation. Hydroxylated protein is then bound by the von Hippel-Lindau tumor suppressor protein (VHL), ubiquitylated, and degraded via proteasome. In contrast, under hypoxia, the hydroxylation reaction is inhibited and HIF-1α is upregulated. HIF-1α then activates the transcription of several genes to mediate a switch from oxidative to glycolytic metabolism and induces angiogenesis by regulating the expression of vascular endothelial growth factor (VEGF) [73, 74]. Koivunen et al. [33] demonstrated that IDH1 mutation attenuates HIF-1α through the activation of HIF prolyl 4-hydroxylase (EGLN), enhancing the proliferation and soft agar growth of NHA.

While several studies demonstrated that the IDH1 mutation induced aerobic glycolysis via HIF-1α activity [59, 75], other group reported that HIF-1α responsive genes, including lactate dehydrogenase (LDHA) were downregulated; silenced LDHA was associated with increased methylation of the LDHA promoter [76]. Another group showed that IDH1 mutation reduces pyruvate flux to lactate and suppresses monocarboxylate transporters MCT1 and MCT4, which mediate lactate transmembrane transport [77]. IDH mutation also alters pyruvate metabolism, including pyruvate dehydrogenase and pyruvate carboxylase enzymes, resulting in anaplerosis of the TCA cycle [78, 79].

Cancer cells are known to depend on reductive carboxylation (RC) of glutamine-derived α-KG for de novo lipogenesis under hypoxia [80]. It is worth noticing that the RC pathway is inhibited by IDH mutation [55]. Under hypoxia, IDH1 mutation upregulated the contribution of glutamine to lipogenesis [81, 56].

Altered amino acids, glutathione, choline derivatives, and tricarboxylic acid (TCA) cycle intermediates were observed in IDH-mutant cells [82, 83]. Glutamate dehydrogenase (GDH)1 and GDH2 were overexpressed in IDH1-mutant tumors, and the orthotopic growth of an IDH1-mutant glioma is inhibited by a double GDH1/2 knockdown [84]. Another group demonstrated that GDH2 was critical for IDH1-mutation induced metabolic alterations and IDH1-mutant glioma growth [85]. The presence of 2-HG also inhibited ATP synthase and mTOR signaling [41].

Importantly, branched-chain amino acid transaminase (BCAT), which catalyzes the α-KG to glutamate conversion, was expressed at lower levels in IDH1-mutant gliomas than in IDH1-wildtype [86, 87]. As a result, the glutamate level was decreased, and cell proliferation and invasiveness were suppressed in IDH-mutant gliomas [87].

Advertisement

7. Role of extensive resection in IDH1-mutant gliomas

There is a huge amount of evidence showing that surgical resection has a pivotal role in survival benefit of glioma patients. Extensive resection is known to prolong survival in low grade glioma and also in GBM (IDH1-wildtype) [88, 89, 90, 91]. In IDH1-mutant gliomas, an MRI study demonstrated that IDH1-mutant tumors were rarely located in high risk areas of the brain and show unilateral patterns of growth, sharp tumor margins, and less contrast enhancement [92, 93]. Indeed, radiographic atlas revealed IDH1-mutant gliomas were frequently located at frontal lobe [94]. A diffusion-tensor imaging study demonstrated that IDH-mutant GBM has a less invasive phenotype than IDH-wildtype GBM [95]. Intriguingly, patients with IDH1-wildtype gliomas had a reduced neurocognitive function and lower performance score than those with IDH1-mutant gliomas [96]. In addition, lesion volume was not associated with neurocognitive function for patients with IDH1-mutant tumors, but associated for those with IDH1-wildtype tumors [96]. Consequently, IDH1-mutant gliomas may be relatively less invasive to the surrounding eloquent area than IDH-wildtype GBM.

In addition, Beiko et al. [97] reported that extensive resection, including nonenhancing area, prolonged survival in IDH1-mutant anaplastic astrocytoma and glioblastoma. They also mentioned, since IDH1-mutant gliomas were predominantly located at frontal lobe, that maximal resection was relatively amenable. Another group independently demonstrated that gross total resection extended survival in grade III IDH1-mutant gliomas without 1p/19q co-deletion [98]. In contrast, survival advantage was controversial in grade II astrocytoma [99, 100]. These results suggest that for IDH1-mutant gliomas, especially grade III astrocytoma, maximal resection should be considered.

Advertisement

8. Prediction of IDH status

To establish IDH status-based treatment strategies, including surgery, advanced preoperative or intraoperative molecular analysis is important. Magnetic resonance spectroscopy (MRS) can be used to detect 2-HG and glutamate changes [101, 102, 103, 104, 105, 106, 107]. A recent MRS study demonstrated that 2-HG peaks rapidly decrease in accordance with tumor regression, whereas they increase with tumor progression in IDH-mutant gliomas [108], suggesting that 2-HG concentration, measured by MRS, may be a reliable approach to evaluate disease states in IDH-mutant gliomas. In addition, several MR techniques, including diffusion tensor imaging and MR methods for determining relative cerebral blood volume, have been proposed to detect mutant IDH1 noninvasively [109, 110, 111]. Moreover, T2-FLAIR mismatch sign was found as a highly specific imaging marker for IDH-mutant astrocytoma [112, 113, 114]. Intraoperative technologies to assess IDH1 mutation have also been established [115, 116, 117]. These advanced technologies may allow the development of tailored surgical strategies for IDH-mutant gliomas. Other group demonstrated that urinary 2-HG is increased in patients with IDH1-mutant gliomas [118]. These findings indicate the possibility of application of indirectly assessed 2-HG as a clinical biomarker.

Advertisement

9. Treatment vulnerability in IDH-mutant gliomas

9.1 Radiotherapy for IDH-mutant gliomas

It has been shown that there is a higher relative sensitivity to radiotherapy and concurrent temozolomide (TMZ) in IDH1-mutant GBM patients than in those with IDH1-wildtype GBM [119], although there is no prospective clinical evidence of radiation therapy to extend survival in glioma patients with IDH1 mutation. As described above, IDH mutation inhibits NADPH and glutamate production, resulting in reduced glutathione levels and increased reactive oxygen species (ROS) [120, 121, 122, 123]. Conversely, radiosensitivity in IDH1-mutant tumors was diminished by IDH1 inhibitor [124]. These findings support selective vulnerability to radiation therapy in IDH-mutant gliomas.

9.2 Chemotherapeutic evidence for IDH-mutant gliomas

9.2.1 Temozolomide

Current standard management of GBM consists of surgical tumor resection, following local radiotherapy with temozolomide treatment [125]. Additionally, adjuvant TMZ prolonged survival in anaplastic astrocytoma [126]. Several studies demonstrated IDH1-mutation as a predictive biomarker for TMZ sensitivity in low grade gliomas and secondary GBM [127, 128].

Cytotoxicity of TMZ is provoked by the formation of O6-methylguanine (O6G)-DNA adducts. O6G-DNA adducts induce DNA strand break and apoptosis through the O6G-thymine-mediated mismatch repair pathway [129, 130]. It has also been established that the activation of DNA repairing pathways, including methylguanine methyltransferase (MGMT) repair enzyme, together with mismatch repair (MMR) system proteins deficiency, such as mutation-induced MSH2 and MSH6, result in drug resistance [131, 132, 133]. MGMT promoter methylation is highly methylated in IDH1-mutant gliomas, particularly oligodendrogliomas, compared with IDH-wildtype [43].

Some preclinical studies demonstrated that forced IDH mutation sensitized cells to chemotherapy by increased ROS [134, 135, 136]. Conversely, forced IDH1 mutation revealed that IDH1 mutation-induced temozolomide (TMZ) resistance and rapid G2 cell cycle arrest through increased RAD-51-mediated homologous recombination (HR) [137, 138]. Importantly, among DNA adducts, O6G represents less than 10%, while the majority of TMZ-induced DNA lesions are N7-methylguanine (60–80%) and N3-methyladenine (10–20%) adducts, which are immediately repaired through poly(ADP-ribose)polymerase (PARP)-dependent base excision repair (BER) [129, 139, 140]. We have recently shown that there are lower steady state NAD+ levels in IDH1-mutant gliomas [141], and that TMZ immediately induces NAD+ consumption through PARP activation-mediated BER in IDH1-mutant gliomas [142]. Besides, Lu et al. [143] reported that the PARP associated DNA repair pathway was extensively compromised in IDH1-mutant cells due to decreased NAD+ availability, thus, cells were sensitive to TMZ, suggesting that deregulated NAD+ metabolism may be related with chemosensitivity. Taken together, these studies show that IDH mutation may increase susceptibility to chemotherapy; however, it remains unclear if IDH mutation itself promotes TMZ sensitivity.

In contrast, TMZ-induced hypermethylation is a critical problem. Long-term TMZ exposure induces MMR inactivation, followed by DNA hypermutation phenotype. Among numerous mutations, gene alterations in RB and AKT-mTOR pathways promoted malignant progression in IDH1-mutant gliomas [27].

9.2.2 Other chemotherapeutic agents

Sulkowski et al. [144] demonstrated that 2-HG inhibits KDM4A and KDM4B, histone demethylases that play a critical role in double strand repair. As a result, IDH1 mutation suppresses HR and induces PARP inhibitor sensitivity. Additionally, IDH1-mutant downregulates the DNA double strand break sensor ATM by altering histone methylation, resulting in impaired DNA repair. As a result, IDH1 mutation causes DNA damage susceptibility to radiation and daunorubicin and reduces self-renewal of hematopoietic stem cells in acute myeloid leukemia [145].

Advertisement

10. Novel therapeutic target in IDH1-mutant tumors

10.1 Specific IDH inhibitor

In 2013, specific inhibitors for IDH1 and IDH2 mutations were discovered [70, 146]. In IDH2-mutant AML cells, an IDH2R140Q inhibitor induced both histone and DNA demethylation [147]. These effects reversed blocked cell differentiation and resulted in cytotoxicity in vitro [146, 147]. It is interesting to note that histone hypermethylation is more rapidly reversed than DNA hypermethylation [147]. In IDH1-mutant AML cells, differentiation and DNA demethylation were also induced by a next generation IDH1 inhibitor [148]. Since the IDH2 mutation is crucial for proliferation and maintenance of leukemia cells [149], an IDH inhibitor may be used as a novel and efficient chemotherapeutic agent against IDH-mutant AML cells. Indeed, clinical trials demonstrated durable response for IDH1/2-mutant refractory AML patients [150, 151].

In IDH1-mutant glioma cells, Rohle et al. [70] reported that a specific IDH1 inhibitor, AGI-5198, blocked 2-HG production, histone demethylation, cell differentiation, and inhibited cell growth in endogenous IDH1-mutant glioma cells. Other group demonstrated that BAY 1436032, a pan inhibitor of IDH1 mutation, promoted mild cytotoxic effects in vivo [152]. In contrast, we established that, even with a long-term IDH1 inhibitor treatment, 2-HG depletion does not induce demethylation of global-DNA and histones, cell differentiation, nor cytotoxicity [141]. Studies using another IDH1 inhibitor also revealed minimal cytotoxicity despite a rapid decrease in 2-HG levels in glioma cells [153, 154]. Similarly, treatment with an IDH1 inhibitor did not contribute to cytotoxicity, and the CpG island methylation status as well as histone trimethylation levels were largely retained in malignant glioma and chondrosarcoma [155, 156]. Intriguingly, in immortalized human astrocytes with an inducible IDH1R132H expression system, a specific IDH1 inhibitor induced demethylation and inhibited tumorigenesis when forced expression was prior or concomitant to inhibitor treatment, but these effects were not observed if the treatment was delayed [157]. These results indicate that 2-HG depletion or blocked mutant IDH1 might be insufficient to control tumor growth and reprogramming of epigenomic alterations in progressed IDH1-mutant gliomas. Indeed, preliminary results indicate that the 6-month progression-free survival of IDH1-mutant glioma, chondrosarcoma, and cholangiocarcinoma is 25, 56, and 43%, respectively, suggesting that the potential of the IDH1 inhibitor may be weaker in IDH1-mutant gliomas than in other cancers [158].

10.2 Other treatment strategies

10.2.1 DNA demethylating agents

In addition to IDH1 inhibitor treatments, other strategies to control IDH1-mutant tumor cells have been proposed. Because the IDH1 mutation promotes proliferation by blocking DNA demethylation, treatment with DNA demethylating agents reverses DNA methylation and inhibits proliferation in IDH1-mutant cells [71, 159]. Intriguingly, treatment with both the DNA demethylating agent 5-azacytidine (5-Aza) and TMZ demonstrated extensively prolonged survival in an IDH1-mutant orthotopic xenograft model [160].

10.2.2 Bcl-2 family inhibitors

Since 2-HG suppresses the activity of cytochrome c oxidase in mitochondrial complex IV, the mitochondrial threshold for apoptosis was decreased after BCL-2 inhibition in IDH1 and IDH2-mutant AML [161]. Similarly, another Bcl-2 family member, the Bcl-xL inhibitor, induced apoptosis in IDH-mutant cells, including endogenous IDH1-mutant glioma cells [162]. Together, inhibition of Bcl-2 family members may be targetable to control growth in IDH-mutant cells.

10.2.3 DNA damaging agents

Because PLK1 activation provokes a rapid bypass through the G2 checkpoint after TMZ treatment in IDH1-mutant tumors, combination treatments with TMZ and a PLK1 inhibitor significantly suppressed tumor growth in an IDH1-mutant in vivo model [138]. In tumors with ATRX mutation-associated alternative lengthening telomeres (ALT), ATR inhibitor is highly sensitive [163], implying that such inhibition may be useful for treatments of IDH1-mutant astrocytic tumors with positive ALT. IDH1 mutation blocked HR, so-called “BRCA ness” phenotype provided specific sensitivity for PARP inhibitor both in vitro and in vivo [144].

10.2.4 DLL-3 targeting therapy

Since Notch ligand DLL-3 is overexpressed in IDH-mutant gliomas, anti-DLL3 antibody-drug conjugate (ADC), rovalpituzumab tesirine (Rova-T), is a potent therapeutic agent for IDH-mutant gliomas [164].

10.2.5 Vaccination therapy

Schumacher et al. [165] reported an immunological approach to control IDH1-mutant cells. They showed that an epitope derived from the IDH1-mutant amino acid sequence is presented in HLA class II molecules of antigen-presenting cells, which elicit a strong immune response via CD4 + T cells. In addition, they showed that constitutive stimulation with synthetic peptides having the IDH1-mutation sequence developed an immune response that eradicated IDH1 mutated tumors in a mouse model with human HLA molecules. Thus, vaccine therapy targeting for IDH1-mutation is expected to develop for future clinical trial [165, 166]. Moreover, IDH1-mutation caused downregulation of leukocyte chemotaxis, resulting in repression of the tumor-associated immune system including immune cells, such as macrophages [167]. Additionally, tumor infiltrating lymphocytes (TILs) and programmed death ligand 1 (PD-L1) were expressed at low levels in IDH1-mutant gliomas [168]. In contrast, Kohanbash et al. [153] demonstrated reduced expression of cytotoxic T lymphocyte-associated genes and IFN-gamma inducible chemokines in IDH1-mutant cells; these results were reversed by specific IDH1 inhibitor. Therefore, combination treatments with vaccine immunotherapy and IDH1 inhibitor result in enhanced toxicity in IDH-mutant tumors.

10.2.6 Target for altered metabolism

IDH1 mutation induced altered metabolism is also expected as a novel therapeutic target. Based on the fact that the main carbon source for α-KG and 2-HG synthesis in IDH1-mutant cells is glutamine from glutaminolysis, a suitable target therapy would be the use of glutaminase (GLS) inhibitor or anti-diabetic drug metformin via the inhibition of mitochondrial complex I in the electron transport system [83, 169, 170, 171]. Since reduced glutamate blocks glutathione synthesis, inhibition of glutaminase specifically sensitizes IDH-mutant glioma cells to oxidative stress and radiation [86].

Mutant IDH1 alters steady state levels of NAD+ through inhibiting NAPRT1, one rate limiting enzyme for NAD+ biosynthesis. Therefore, inhibition of nicotinamide phosphoribosyltransferase (NAMPT), another rate limiting enzyme, induced high cytotoxicity in IDH1-mutant patient-derived glioma cells [141]. Since TMZ rapidly consumes NAD+ through PARP activation, combination treatments with TMZ and NAMPT inhibitor further enhanced NAD+ depletion-mediated cytotoxicity in IDH1-mutant cancers [142]. Similarly, Lu et al. [143] reported that the PARP-associated DNA repair pathway was extensively compromised in IDH1-mutant cells due to decreased NAD+ availability, thus sensitive to TMZ.

Because of the relationships between IDH1 mutation and MYC activation [38, 40, 172], target therapy to regulate MYC, by using bromodomain and extra-terminal (BET) inhibitors, CDK7 or MYC-induced glycolysis may be used for IDH-mutant gliomas [40, 173, 174, 175]. Given the results of these studies, IDH1 mutation-specific biological alterations and metabolic feature may be expected as novel therapeutic targets.

11. Conclusions

In summary, investigations on IDH mutations enabled distinctive tumor classification and may allow the development of specific therapeutic strategies. Further preclinical and clinical studies are warranted to overcome the outcomes of cancer development in IDH-mutant glioma patients.

References

  1. 1. Sjoblom T, Jones S, Wood LD, Parsons DW, Lin J, Barber TD, et al. The consensus coding sequences of human breast and colorectal cancers. Science. 2006;314(5797):268-274
  2. 2. Parsons DW, Jones S, Zhang X, Lin JC, Leary RJ, Angenendt P, et al. An integrated genomic analysis of human glioblastoma multiforme. Science. 2008;321(5897):1807-1812
  3. 3. Ichimura K, Pearson DM, Kocialkowski S, Backlund LM, Chan R, Jones DT, et al. IDH1 mutations are present in the majority of common adult gliomas but rare in primary glioblastomas. Neuro-Oncology. 2009;11(4):341-347
  4. 4. Yan H, Parsons DW, Jin G, McLendon R, Rasheed BA, Yuan W, et al. IDH1 and IDH2 mutations in gliomas. The New England Journal of Medicine. 2009;360(8):765-773
  5. 5. Balss J, Meyer J, Mueller W, Korshunov A, Hartmann C, von Deimling A. Analysis of the IDH1 codon 132 mutation in brain tumors. Acta Neuropathologica. 2008;116(6):597-602
  6. 6. Pusch S, Schweizer L, Beck AC, Lehmler JM, Weissert S, Balss J, et al. D-2-hydroxyglutarate producing neo-enzymatic activity inversely correlates with frequency of the type of isocitrate dehydrogenase 1 mutations found in glioma. Acta Neuropathologica Communications. 2014;2:19
  7. 7. Hartmann C, Meyer J, Balss J, Capper D, Mueller W, Christians A, et al. Type and frequency of IDH1 and IDH2 mutations are related to astrocytic and oligodendroglial differentiation and age: A study of 1,010 diffuse gliomas. Acta Neuropathologica. 2009;118(4):469-474
  8. 8. Mardis ER, Ding L, Dooling DJ, Larson DE, McLellan MD, Chen K, et al. Recurring mutations found by sequencing an acute myeloid leukemia genome. The New England Journal of Medicine. 2009;361(11):1058-1066
  9. 9. Marcucci G, Maharry K, Wu YZ, Radmacher MD, Mrozek K, Margeson D, et al. IDH1 and IDH2 gene mutations identify novel molecular subsets within de novo cytogenetically normal acute myeloid leukemia: A Cancer and Leukemia Group B study. Journal of Clinical Oncology. 2010;28(14):2348-2355
  10. 10. Cairns RA, Iqbal J, Lemonnier F, Kucuk C, de Leval L, Jais JP, et al. IDH2 mutations are frequent in angioimmunoblastic T-cell lymphoma. Blood. 2012;119(8):1901-1903
  11. 11. Amary MF, Bacsi K, Maggiani F, Damato S, Halai D, Berisha F, et al. IDH1 and IDH2 mutations are frequent events in central chondrosarcoma and central and periosteal chondromas but not in other mesenchymal tumours. The Journal of Pathology. 2011;224(3):334-343
  12. 12. Kipp BR, Voss JS, Kerr SE, Barr Fritcher EG, Graham RP, Zhang L, et al. Isocitrate dehydrogenase 1 and 2 mutations in cholangiocarcinoma. Human Pathology. 2012;43(10):1552-1558
  13. 13. Wang P, Dong Q , Zhang C, Kuan PF, Liu Y, Jeck WR, et al. Mutations in isocitrate dehydrogenase 1 and 2 occur frequently in intrahepatic cholangiocarcinomas and share hypermethylation targets with glioblastomas. Oncogene. 2013;32(25):3091-3100
  14. 14. Kang MR, Kim MS, Oh JE, Kim YR, Song SY, Seo SI, et al. Mutational analysis of IDH1 codon 132 in glioblastomas and other common cancers. International Journal of Cancer. 2009;125(2):353-355
  15. 15. Fathi AT, Sadrzadeh H, Comander AH, Higgins MJ, Bardia A, Perry A, et al. Isocitrate dehydrogenase 1 (IDH1) mutation in breast adenocarcinoma is associated with elevated levels of serum and urine 2-hydroxyglutarate. The Oncologist. 2014;19(6):602-607
  16. 16. Li-Chang HH, Kasaian K, Ng Y, Lum A, Kong E, Lim H, et al. Retrospective review using targeted deep sequencing reveals mutational differences between gastroesophageal junction and gastric carcinomas. BMC Cancer. 2015;15:32
  17. 17. Nonoguchi N, Ohta T, Oh JE, Kim YH, Kleihues P, Ohgaki H. TERT promoter mutations in primary and secondary glioblastomas. Acta Neuropathologica. 2013;126(6):931-937
  18. 18. Ohgaki H, Kleihues P. The definition of primary and secondary glioblastoma. Clinical Cancer Research. 2013;19(4):764-772
  19. 19. Louis DN, Perry A, Reifenberger G, von Deimling A, Figarella-Branger D, Cavenee WK, et al. The 2016 world health organization classification of tumors of the central nervous system: A summary. Acta Neuropathologica. 2016;131(6):803-820
  20. 20. Jiao Y, Killela PJ, Reitman ZJ, Rasheed AB, Heaphy CM, de Wilde RF, et al. Frequent ATRX, CIC, FUBP1 and IDH1 mutations refine the classification of malignant gliomas. Oncotarget. 2012;3(7):709-722
  21. 21. Kannan K, Inagaki A, Silber J, Gorovets D, Zhang J, Kastenhuber ER, et al. Whole-exome sequencing identifies ATRX mutation as a key molecular determinant in lower-grade glioma. Oncotarget. 2012;3(10):1194-1203
  22. 22. Liu XY, Gerges N, Korshunov A, Sabha N, Khuong-Quang DA, Fontebasso AM, et al. Frequent ATRX mutations and loss of expression in adult diffuse astrocytic tumors carrying IDH1/IDH2 and TP53 mutations. Acta Neuropathologica. 2012;124(5):615-625
  23. 23. Arita H, Narita Y, Fukushima S, Tateishi K, Matsushita Y, Yoshida A, et al. Upregulating mutations in the TERT promoter commonly occur in adult malignant gliomas and are strongly associated with total 1p19q loss. Acta Neuropathologica. 2013;126(2):267-276
  24. 24. Killela PJ, Reitman ZJ, Jiao Y, Bettegowda C, Agrawal N, Diaz LA Jr, et al. TERT promoter mutations occur frequently in gliomas and a subset of tumors derived from cells with low rates of self-renewal. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(15):6021-6026
  25. 25. Watanabe T, Nobusawa S, Kleihues P, Ohgaki H. IDH1 mutations are early events in the development of astrocytomas and oligodendrogliomas. The American Journal of Pathology. 2009;174(4):1149-1153
  26. 26. Juratli TA, Kirsch M, Robel K, Soucek S, Geiger K, von Kummer R, et al. IDH mutations as an early and consistent marker in low-grade astrocytomas WHO grade II and their consecutive secondary high-grade gliomas. Journal of Neuro-Oncology. 2012;108(3):403-410
  27. 27. Johnson BE, Mazor T, Hong C, Barnes M, Aihara K, McLean CY, et al. Mutational analysis reveals the origin and therapy-driven evolution of recurrent glioma. Science. 2014;343(6167):189-193
  28. 28. Suzuki H, Aoki K, Chiba K, Sato Y, Shiozawa Y, Shiraishi Y, et al. Mutational landscape and clonal architecture in grade II and III gliomas. Nature Genetics. 2015;47(5):458-468
  29. 29. Mazor T, Chesnelong C, Pankov A, Jalbert LE, Hong C, Hayes J, et al. Clonal expansion and epigenetic reprogramming following deletion or amplification of mutant IDH1. Proceedings of the National Academy of Sciences of the United States of America. 2017;114(40):10743-10748
  30. 30. Ohba S, Mukherjee J, Johannessen TC, Mancini A, Chow TT, Wood M, et al. Mutant IDH1 expression drives TERT promoter reactivation as part of the cellular transformation process. Cancer Research. 2016;76(22):6680-6689
  31. 31. Tirosh I, Venteicher AS, Hebert C, Escalante LE, Patel AP, Yizhak K, et al. Single-cell RNA-seq supports a developmental hierarchy in human oligodendroglioma. Nature. 2016;539(7628):309-313
  32. 32. Venteicher AS, Tirosh I, Hebert C, Yizhak K, Neftel C, Filbin MG, et al. Decoupling genetics, lineages, and microenvironment in IDH-mutant gliomas by single-cell RNA-seq. Science. 2017;355(6332)
  33. 33. Koivunen P, Lee S, Duncan CG, Lopez G, Lu G, Ramkissoon S, et al. Transformation by the (R)-enantiomer of 2-hydroxyglutarate linked to EGLN activation. Nature. 2012;483(7390):484-488
  34. 34. Piaskowski S, Bienkowski M, Stoczynska-Fidelus E, Stawski R, Sieruta M, Szybka M, et al. Glioma cells showing IDH1 mutation cannot be propagated in standard cell culture conditions. British Journal of Cancer. 2011;104(6):968-970
  35. 35. Luchman HA, Stechishin OD, Dang NH, Blough MD, Chesnelong C, Kelly JJ, et al. An in vivo patient-derived model of endogenous IDH1-mutant glioma. Neuro-Oncology. 2012;14(2):184-191
  36. 36. Balvers RK, Kleijn A, Kloezeman JJ, French PJ, Kremer A, van den Bent MJ, et al. Serum-free culture success of glial tumors is related to specific molecular profiles and expression of extracellular matrix-associated gene modules. Neuro-Oncology. 2013;15(12):1684-1695
  37. 37. Bardella C, Al-Dalahmah O, Krell D, Brazauskas P, Al-Qahtani K, Tomkova M, et al. Expression of Idh1R132H in the murine subventricular zone stem cell niche recapitulates features of early gliomagenesis. Cancer Cell. 2016;30(4):578-594
  38. 38. Wakimoto H, Tanaka S, Curry WT, Loebel F, Zhao D, Tateishi K, et al. Targetable signaling pathway mutations are associated with malignant phenotype in IDH-mutant gliomas. Clinical Cancer Research. 2014;20(11):2898-2909
  39. 39. Kamoun A, Idbaih A, Dehais C, Elarouci N, Carpentier C, Letouze E, et al. Integrated multi-omics analysis of oligodendroglial tumours identifies three subgroups of 1p/19q co-deleted gliomas. Nature Communications. 2016;7:11263
  40. 40. Bai H, Harmanci AS, Erson-Omay EZ, Li J, Coskun S, Simon M, et al. Integrated genomic characterization of IDH1-mutant glioma malignant progression. Nature Genetics. 2016;48(1):59-66
  41. 41. Reuss DE, Mamatjan Y, Schrimpf D, Capper D, Hovestadt V, Kratz A, et al. IDH mutant diffuse and anaplastic astrocytomas have similar age at presentation and little difference in survival: A grading problem for WHO. Acta Neuropathologica. 2015;129(6):867-873
  42. 42. Nobusawa S, Watanabe T, Kleihues P, Ohgaki H. IDH1 mutations as molecular signature and predictive factor of secondary glioblastomas. Clinical Cancer Research. 2009;15(19):6002-6007
  43. 43. Sanson M, Marie Y, Paris S, Idbaih A, Laffaire J, Ducray F, et al. Isocitrate dehydrogenase 1 codon 132 mutation is an important prognostic biomarker in gliomas. Journal of Clinical Oncology. 2009;27(25):4150-4154
  44. 44. Wick W, Hartmann C, Engel C, Stoffels M, Felsberg J, Stockhammer F, et al. NOA-04 randomized phase III trial of sequential radiochemotherapy of anaplastic glioma with procarbazine, lomustine, and vincristine or temozolomide. Journal of Clinical Oncology. 2009;27(35):5874-5880
  45. 45. Killela PJ, Pirozzi CJ, Healy P, Reitman ZJ, Lipp E, Rasheed BA, et al. Mutations in IDH1, IDH2, and in the TERT promoter define clinically distinct subgroups of adult malignant gliomas. Oncotarget. 2014;5(6):1515-1525
  46. 46. Minniti G, Scaringi C, Arcella A, Lanzetta G, Di Stefano D, Scarpino S, et al. IDH1 mutation and MGMT methylation status predict survival in patients with anaplastic astrocytoma treated with temozolomide-based chemoradiotherapy. Journal of Neuro-Oncology. 2014;118(2):377-383
  47. 47. Eckel-Passow JE, Lachance DH, Molinaro AM, Walsh KM, Decker PA, Sicotte H, et al. Glioma groups based on 1p/19q, IDH, and TERT promoter mutations in tumors. The New England Journal of Medicine. 2015;372(26):2499-2508
  48. 48. Hartmann C, Hentschel B, Wick W, Capper D, Felsberg J, Simon M, et al. Patients with IDH1 wild type anaplastic astrocytomas exhibit worse prognosis than IDH1-mutated glioblastomas, and IDH1 mutation status accounts for the unfavorable prognostic effect of higher age: Implications for classification of gliomas. Acta Neuropathologica. 2010;120(6):707-718
  49. 49. Pekmezci M, Rice T, Molinaro AM, Walsh KM, Decker PA, Hansen H, et al. Adult infiltrating gliomas with WHO 2016 integrated diagnosis: Additional prognostic roles of ATRX and TERT. Acta Neuropathologica. 2017;133(6):1001-1016
  50. 50. Arita H, Yamasaki K, Matsushita Y, Nakamura T, Shimokawa A, Takami H, et al. A combination of TERT promoter mutation and MGMT methylation status predicts clinically relevant subgroups of newly diagnosed glioblastomas. Acta Neuropathologica Communications. 2016;4(1):79
  51. 51. Shirahata M, Ono T, Stichel D, Schrimpf D, Reuss DE, Sahm F, et al. Novel, improved grading system(s) for IDH-mutant astrocytic gliomas. Acta Neuropathologica. 2018;136(1):153-166
  52. 52. Halani SH, Yousefi S, Vega JV, Rossi MR, Zhao Z, Amrollahi F, et al. Multi-faceted computational assessment of risk and progression in oligodendroglioma implicates NOTCH and PI3K pathways. NPJ Precision Oncology. 2018;2:24
  53. 53. Aoki K, Nakamura H, Suzuki H, Matsuo K, Kataoka K, Shimamura T, et al. Prognostic relevance of genetic alterations in diffuse lower-grade gliomas. Neuro-Oncology. 2018;20(1):66-77
  54. 54. Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature. 2009;462(7274):739-744
  55. 55. Leonardi R, Subramanian C, Jackowski S, Rock CO. Cancer-associated isocitrate dehydrogenase mutations inactivate NADPH-dependent reductive carboxylation. The Journal of Biological Chemistry. 2012;287(18):14615-14620
  56. 56. Grassian AR, Parker SJ, Davidson SM, Divakaruni AS, Green CR, Zhang X, et al. IDH1 mutations alter citric acid cycle metabolism and increase dependence on oxidative mitochondrial metabolism. Cancer Research. 2014;74(12):3317-3331
  57. 57. Ward PS, Patel J, Wise DR, Abdel-Wahab O, Bennett BD, Coller HA, et al. The common feature of leukemia-associated IDH1 and IDH2 mutations is a neomorphic enzyme activity converting alpha-ketoglutarate to 2-hydroxyglutarate. Cancer Cell. 2010;17(3):225-234
  58. 58. Jin G, Reitman ZJ, Duncan CG, Spasojevic I, Gooden DM, Rasheed BA, et al. Disruption of wild-type IDH1 suppresses D-2-hydroxyglutarate production in IDH1-mutated gliomas. Cancer Research. 2013;73(2):496-501
  59. 59. Zhao S, Lin Y, Xu W, Jiang W, Zha Z, Wang P, et al. Glioma-derived mutations in IDH1 dominantly inhibit IDH1 catalytic activity and induce HIF-1alpha. Science. 2009;324(5924):261-265
  60. 60. Xu W, Yang H, Liu Y, Yang Y, Wang P, Kim SH, et al. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of alpha-ketoglutarate-dependent dioxygenases. Cancer Cell. 2011;19(1):17-30
  61. 61. Figueroa ME, Abdel-Wahab O, Lu C, Ward PS, Patel J, Shih A, et al. Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell. 2010;18(6):553-567
  62. 62. Noushmehr H, Weisenberger DJ, Diefes K, Phillips HS, Pujara K, Berman BP, et al. Identification of a CpG island methylator phenotype that defines a distinct subgroup of glioma. Cancer Cell. 2010;17(5):510-522
  63. 63. Sasaki M, Knobbe CB, Munger JC, Lind EF, Brenner D, Brustle A, et al. IDH1(R132H) mutation increases murine haematopoietic progenitors and alters epigenetics. Nature. 2012;488(7413):656-659
  64. 64. Turcan S, Rohle D, Goenka A, Walsh LA, Fang F, Yilmaz E, et al. IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature. 2012;483(7390):479-483
  65. 65. de Souza CF, Sabedot TS, Malta TM, Stetson L, Morozova O, Sokolov A, et al. A distinct DNA methylation shift in a subset of glioma CpG island methylator phenotypes during tumor recurrence. Cell Reports. 2018;23(2):637-651
  66. 66. Verhaak RG, Hoadley KA, Purdom E, Wang V, Qi Y, Wilkerson MD, et al. Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell. 2010;17(1):98-110
  67. 67. Flavahan WA, Drier Y, Liau BB, Gillespie SM, Venteicher AS, Stemmer-Rachamimov AO, et al. Insulator dysfunction and oncogene activation in IDH mutant gliomas. Nature. 2016;529(7584):110-114
  68. 68. Chowdhury R, Yeoh KK, Tian YM, Hillringhaus L, Bagg EA, Rose NR, et al. The oncometabolite 2-hydroxyglutarate inhibits histone lysine demethylases. EMBO Reports. 2011;12(5):463-469
  69. 69. Lu C, Ward PS, Kapoor GS, Rohle D, Turcan S, Abdel-Wahab O, et al. IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature. 2012;483(7390):474-478
  70. 70. Rohle D, Popovici-Muller J, Palaskas N, Turcan S, Grommes C, Campos C, et al. An inhibitor of mutant IDH1 delays growth and promotes differentiation of glioma cells. Science. 2013;340(6132):626-630
  71. 71. Turcan S, Fabius AW, Borodovsky A, Pedraza A, Brennan C, Huse J, et al. Efficient induction of differentiation and growth inhibition in IDH1 mutant glioma cells by the DNMT inhibitor decitabine. Oncotarget. 2013;4(10):1729-1736
  72. 72. Sasaki M, Knobbe CB, Itsumi M, Elia AJ, Harris IS, Chio II, et al. D-2-hydroxyglutarate produced by mutant IDH1 perturbs collagen maturation and basement membrane function. Genes & Development. 2012;26(18):2038-2049
  73. 73. Semenza GL. HIF-1 mediates metabolic responses to intratumoral hypoxia and oncogenic mutations. The Journal of Clinical Investigation. 2013;123(9):3664-3671
  74. 74. Semenza GL. Regulation of metabolism by hypoxia-inducible factor 1. Cold Spring Harbor Symposia on Quantitative Biology. 2011;76:347-353
  75. 75. Nie Q , Guo P, Guo L, Lan J, Lin Y, Guo F, et al. Overexpression of isocitrate dehydrogenase-1R(1)(3)(2)H enhances the proliferation of A172 glioma cells via aerobic glycolysis. Molecular Medicine Reports. 2015;11(5):3715-3721
  76. 76. Chesnelong C, Chaumeil MM, Blough MD, Al-Najjar M, Stechishin OD, Chan JA, et al. Lactate dehydrogenase A silencing in IDH mutant gliomas. Neuro-Oncology. 2014;16(5):686-695
  77. 77. Viswanath P, Najac C, Izquierdo-Garcia JL, Pankov A, Hong C, Eriksson P, et al. Mutant IDH1 expression is associated with down-regulation of monocarboxylate transporters. Oncotarget. 2016;7(23):34942-34955
  78. 78. Izquierdo-Garcia JL, Cai LM, Chaumeil MM, Eriksson P, Robinson AE, Pieper RO, et al. Glioma cells with the IDH1 mutation modulate metabolic fractional flux through pyruvate carboxylase. PLoS ONE. 2014;9(9):e108289
  79. 79. Izquierdo-Garcia JL, Viswanath P, Eriksson P, Cai L, Radoul M, Chaumeil MM, et al. IDH1 mutation induces reprogramming of pyruvate metabolism. Cancer Research. 2015;75(15):2999-3009
  80. 80. Metallo CM, Gameiro PA, Bell EL, Mattaini KR, Yang J, Hiller K, et al. Reductive glutamine metabolism by IDH1 mediates lipogenesis under hypoxia. Nature. 2012;481(7381):380-384
  81. 81. Reitman ZJ, Duncan CG, Poteet E, Winters A, Yan LJ, Gooden DM, et al. Cancer-associated isocitrate dehydrogenase 1 (IDH1) R132H mutation and d-2-hydroxyglutarate stimulate glutamine metabolism under hypoxia. The Journal of Biological Chemistry. 2014;289(34):23318-23328
  82. 82. Reitman ZJ, Jin G, Karoly ED, Spasojevic I, Yang J, Kinzler KW, et al. Profiling the effects of isocitrate dehydrogenase 1 and 2 mutations on the cellular metabolome. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(8):3270-3275
  83. 83. Ohka F, Ito M, Ranjit M, Senga T, Motomura A, Motomura K, et al. Quantitative metabolome analysis profiles activation of glutaminolysis in glioma with IDH1 mutation. Tumour Biology. 2014;35(6):5911-5920
  84. 84. Chen R, Nishimura MC, Kharbanda S, Peale F, Deng Y, Daemen A, et al. Hominoid-specific enzyme GLUD2 promotes growth of IDH1R132H glioma. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(39):14217-14222
  85. 85. Waitkus MS, Pirozzi CJ, Moure CJ, Diplas BH, Hansen LJ, Carpenter AB, et al. Adaptive evolution of the GDH2 allosteric domain promotes gliomagenesis by resolving IDH1(R132H)-induced metabolic liabilities. Cancer Research. 2018;78(1):36-50
  86. 86. McBrayer SK, Mayers JR, DiNatale GJ, Shi DD, Khanal J, Chakraborty AA, et al. Transaminase inhibition by 2-hydroxyglutarate impairs glutamate biosynthesis and redox homeostasis in glioma. Cell. 2018;175(1):101-116 e25
  87. 87. Tonjes M, Barbus S, Park YJ, Wang W, Schlotter M, Lindroth AM, et al. BCAT1 promotes cell proliferation through amino acid catabolism in gliomas carrying wild-type IDH1. Nature Medicine. 2013;19(7):901-908
  88. 88. Lacroix M, Abi-Said D, Fourney DR, Gokaslan ZL, Shi W, DeMonte F, et al. A multivariate analysis of 416 patients with glioblastoma multiforme: Prognosis, extent of resection, and survival. Journal of Neurosurgery. 2001;95(2):190-198
  89. 89. Smith JS, Chang EF, Lamborn KR, Chang SM, Prados MD, Cha S, et al. Role of extent of resection in the long-term outcome of low-grade hemispheric gliomas. Journal of Clinical Oncology. 2008;26(8):1338-1345
  90. 90. Sanai N, Polley MY, McDermott MW, Parsa AT, Berger MS. An extent of resection threshold for newly diagnosed glioblastomas. Journal of Neurosurgery. 2011;115(1):3-8
  91. 91. Marko NF, Weil RJ, Schroeder JL, Lang FF, Suki D, Sawaya RE. Extent of resection of glioblastoma revisited: Personalized survival modeling facilitates more accurate survival prediction and supports a maximum-safe-resection approach to surgery. Journal of Clinical Oncology. 2014;32(8):774-782
  92. 92. Qi S, Yu L, Li H, Ou Y, Qiu X, Ding Y, et al. Isocitrate dehydrogenase mutation is associated with tumor location and magnetic resonance imaging characteristics in astrocytic neoplasms. Oncology Letters. 2014;7(6):1895-1902
  93. 93. Darlix A, Deverdun J, Menjot de Champfleur N, Castan F, Zouaoui S, Rigau V, et al. IDH mutation and 1p19q codeletion distinguish two radiological patterns of diffuse low-grade gliomas. Journal of Neuro-Oncology. 2017;133(1):37-45
  94. 94. Ellingson BM, Lai A, Harris RJ, Selfridge JM, Yong WH, Das K, et al. Probabilistic radiographic atlas of glioblastoma phenotypes. AJNR. American Journal of Neuroradiology. 2013;34(3):533-540
  95. 95. Price SJ, Allinson K, Liu H, Boonzaier NR, Yan JL, Lupson VC, et al. Less invasive phenotype found in isocitrate dehydrogenase-mutated glioblastomas than in isocitrate dehydrogenase wild-type glioblastomas: A diffusion-tensor imaging study. Radiology. 2017;283(1):215-221
  96. 96. Wefel JS, Noll KR, Rao G, Cahill DP. Neurocognitive function varies by IDH1 genetic mutation status in patients with malignant glioma prior to surgical resection. Neuro-Oncology. 2016;18(12):1656-1663
  97. 97. Beiko J, Suki D, Hess KR, Fox BD, Cheung V, Cabral M, et al. IDH1 mutant malignant astrocytomas are more amenable to surgical resection and have a survival benefit associated with maximal surgical resection. Neuro-Oncology. 2014;16(1):81-91
  98. 98. Kawaguchi T, Sonoda Y, Shibahara I, Saito R, Kanamori M, Kumabe T, et al. Impact of gross total resection in patients with WHO grade III glioma harboring the IDH 1/2 mutation without the 1p/19q co-deletion. Journal of Neuro-Oncology. 2016;129(3):505-514
  99. 99. Wijnenga MMJ, French PJ, Dubbink HJ, Dinjens WNM, Atmodimedjo PN, Kros JM, et al. The impact of surgery in molecularly defined low-grade glioma: An integrated clinical, radiological, and molecular analysis. Neuro-Oncology. 2018;20(1):103-112
  100. 100. Patel T, Bander ED, Venn RA, Powell T, Cederquist GY, Schaefer PM, et al. The role of extent of resection in IDH1 wild-type or mutant low-grade gliomas. Neurosurgery. 2018;82(6):808-814
  101. 101. Andronesi OC, Kim GS, Gerstner E, Batchelor T, Tzika AA, Fantin VR, et al. Detection of 2-hydroxyglutarate in IDH-mutated glioma patients by in vivo spectral-editing and 2D correlation magnetic resonance spectroscopy. Science Translational Medicine. 2012;4(116):116ra4
  102. 102. Choi C, Ganji SK, DeBerardinis RJ, Hatanpaa KJ, Rakheja D, Kovacs Z, et al. 2-hydroxyglutarate detection by magnetic resonance spectroscopy in IDH-mutated patients with gliomas. Nature Medicine. 2012;18(4):624-629
  103. 103. Lazovic J, Soto H, Piccioni D, Lou JR, Li S, Mirsadraei L, et al. Detection of 2-hydroxyglutaric acid in vivo by proton magnetic resonance spectroscopy in U87 glioma cells overexpressing isocitrate dehydrogenase-1 mutation. Neuro-Oncology. 2012;14(12):1465-1472
  104. 104. Andronesi OC, Rapalino O, Gerstner E, Chi A, Batchelor TT, Cahill DP, et al. Detection of oncogenic IDH1 mutations using magnetic resonance spectroscopy of 2-hydroxyglutarate. The Journal of Clinical Investigation. 2013;123(9):3659-3663
  105. 105. de la Fuente MI, Young RJ, Rubel J, Rosenblum M, Tisnado J, Briggs S, et al. Integration of 2-hydroxyglutarate-proton magnetic resonance spectroscopy into clinical practice for disease monitoring in isocitrate dehydrogenase-mutant glioma. Neuro-Oncology. 2016;18(2):283-290
  106. 106. Emir UE, Larkin SJ, de Pennington N, Voets N, Plaha P, Stacey R, et al. Noninvasive quantification of 2-hydroxyglutarate in human gliomas with IDH1 and IDH2 mutations. Cancer Research. 2016;76(1):43-49
  107. 107. Nagashima H, Tanaka K, Sasayama T, Irino Y, Sato N, Takeuchi Y, et al. Diagnostic value of glutamate with 2-hydroxyglutarate in magnetic resonance spectroscopy for IDH1 mutant glioma. Neuro-Oncology. 2016;18(11):1559-1568
  108. 108. Choi C, Raisanen JM, Ganji SK, Zhang S, McNeil SS, An Z, et al. Prospective longitudinal analysis of 2-hydroxyglutarate magnetic resonance spectroscopy identifies broad clinical utility for the management of patients with IDH-mutant glioma. Journal of Clinical Oncology. 2016;34(33):4030-4039
  109. 109. Tan WL, Huang WY, Yin B, Xiong J, Wu JS, Geng DY. Can diffusion tensor imaging noninvasively detect IDH1 gene mutations in astrogliomas? A retrospective study of 112 cases. AJNR. American Journal of Neuroradiology. 2014;35(5):920-927
  110. 110. Kickingereder P, Sahm F, Radbruch A, Wick W, Heiland S, Deimling A, et al. IDH mutation status is associated with a distinct hypoxia/angiogenesis transcriptome signature which is non-invasively predictable with rCBV imaging in human glioma. Scientific Reports. 2015;5:16238
  111. 111. Yamashita K, Hiwatashi A, Togao O, Kikuchi K, Hatae R, Yoshimoto K, et al. MR imaging-based analysis of glioblastoma multiforme: Estimation of IDH1 mutation status. AJNR. American Journal of Neuroradiology. 2016;37(1):58-65
  112. 112. Juratli TA, Tummala SS, Riedl A, Daubner D, Hennig S, Penson T, et al. Radiographic assessment of contrast enhancement and T2/FLAIR mismatch sign in lower grade gliomas: Correlation with molecular groups. Journal of Neuro-Oncology. 2019;141(2):327-335
  113. 113. Broen MPG, Smits M, Wijnenga MMJ, Dubbink HJ, Anten M, Schijns O, et al. The T2-FLAIR mismatch sign as an imaging marker for non-enhancing IDH-mutant, 1p/19q-intact lower-grade glioma: A validation study. Neuro-Oncology. 2018;20(10):1393-1399
  114. 114. Patel SH, Poisson LM, Brat DJ, Zhou Y, Cooper L, Snuderl M, et al. T2-FLAIR mismatch, an imaging biomarker for IDH and 1p/19q status in lower-grade gliomas: A TCGA/TCIA project. Clinical Cancer Research. 2017;23(20):6078-6085
  115. 115. Kanamori M, Kikuchi A, Watanabe M, Shibahara I, Saito R, Yamashita Y, et al. Rapid and sensitive intraoperative detection of mutations in the isocitrate dehydrogenase 1 and 2 genes during surgery for glioma. Journal of Neurosurgery. 2014;120(6):1288-1297
  116. 116. Santagata S, Eberlin LS, Norton I, Calligaris D, Feldman DR, Ide JL, et al. Intraoperative mass spectrometry mapping of an onco-metabolite to guide brain tumor surgery. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(30):11121-11126
  117. 117. Shankar GM, Francis JM, Rinne ML, Ramkissoon SH, Huang FW, Venteicher AS, et al. Rapid intraoperative molecular characterization of glioma. JAMA Oncology. 2015;1(5):662-667
  118. 118. Fathi AT, Nahed BV, Wander SA, Iafrate AJ, Borger DR, Hu R, et al. Elevation of urinary 2-hydroxyglutarate in IDH-mutant glioma. The Oncologist. 2016;21(2):214-219
  119. 119. Tran AN, Lai A, Li S, Pope WB, Teixeira S, Harris RJ, et al. Increased sensitivity to radiochemotherapy in IDH1 mutant glioblastoma as demonstrated by serial quantitative MR volumetry. Neuro-Oncology. 2014;16(3):414-420
  120. 120. Li S, Chou AP, Chen W, Chen R, Deng Y, Phillips HS, et al. Overexpression of isocitrate dehydrogenase mutant proteins renders glioma cells more sensitive to radiation. Neuro-Oncology. 2013;15(1):57-68
  121. 121. Wang XW, Labussiere M, Valable S, Peres EA, Guillamo JS, Bernaudin M, et al. IDH1(R132H) mutation increases U87 glioma cell sensitivity to radiation therapy in hypoxia. BioMed Research International. 2014;2014:198697
  122. 122. Kessler J, Guttler A, Wichmann H, Rot S, Kappler M, Bache M, et al. IDH1(R132H) mutation causes a less aggressive phenotype and radiosensitizes human malignant glioma cells independent of the oxygenation status. Radiotherapy and Oncology. 2015;116(3):381-387
  123. 123. Bleeker FE, Atai NA, Lamba S, Jonker A, Rijkeboer D, Bosch KS, et al. The prognostic IDH1( R132 ) mutation is associated with reduced NADP+-dependent IDH activity in glioblastoma. Acta Neuropathologica. 2010;119(4):487-494
  124. 124. Molenaar RJ, Botman D, Smits MA, Hira VV, van Lith SA, Stap J, et al. Radioprotection of IDH1-mutated cancer cells by the IDH1-mutant inhibitor AGI-5198. Cancer Research. 2015;75(22):4790-4802
  125. 125. Stupp R, Mason WP, van den Bent MJ, Weller M, Fisher B, Taphoorn MJ, et al. Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. The New England Journal of Medicine. 2005;352(10):987-996
  126. 126. van den Bent MJ, Baumert B, Erridge SC, Vogelbaum MA, Nowak AK, Sanson M, et al. Interim results from the CATNON trial (EORTC study 26053-22054) of treatment with concurrent and adjuvant temozolomide for 1p/19q non-co-deleted anaplastic glioma: A phase 3, randomised, open-label intergroup study. Lancet. 2017;390(10103):1645-1653
  127. 127. Houillier C, Wang X, Kaloshi G, Mokhtari K, Guillevin R, Laffaire J, et al. IDH1 or IDH2 mutations predict longer survival and response to temozolomide in low-grade gliomas. Neurology. 2010;75(17):1560-1566
  128. 128. SongTao Q , Lei Y, Si G, YanQing D, HuiXia H, XueLin Z, et al. IDH mutations predict longer survival and response to temozolomide in secondary glioblastoma. Cancer Science. 2012;103(2):269-273
  129. 129. Fu D, Calvo JA, Samson LD. Balancing repair and tolerance of DNA damage caused by alkylating agents. Nature Reviews. Cancer. 2012;12(2):104-120
  130. 130. Fan CH, Liu WL, Cao H, Wen C, Chen L, Jiang G. O6-methylguanine DNA methyltransferase as a promising target for the treatment of temozolomide-resistant gliomas. Cell Death & Disease. 2013;4:e876
  131. 131. Hegi ME, Diserens AC, Gorlia T, Hamou MF, de Tribolet N, Weller M, et al. MGMT gene silencing and benefit from temozolomide in glioblastoma. The New England Journal of Medicine. 2005;352(10):997-1003
  132. 132. Hunter C, Smith R, Cahill DP, Stephens P, Stevens C, Teague J, et al. A hypermutation phenotype and somatic MSH6 mutations in recurrent human malignant gliomas after alkylator chemotherapy. Cancer Research. 2006;66(8):3987-3991
  133. 133. Cahill DP, Levine KK, Betensky RA, Codd PJ, Romany CA, Reavie LB, et al. Loss of the mismatch repair protein MSH6 in human glioblastomas is associated with tumor progression during temozolomide treatment. Clinical Cancer Research. 2007;13(7):2038-2045
  134. 134. Fu Y, Huang R, Zheng Y, Zhang Z, Liang A. Glioma-derived mutations in isocitrate dehydrogenase 2 beneficial to traditional chemotherapy. Biochemical and Biophysical Research Communications. 2011;410(2):218-223
  135. 135. Mohrenz IV, Antonietti P, Pusch S, Capper D, Balss J, Voigt S, et al. Isocitrate dehydrogenase 1 mutant R132H sensitizes glioma cells to BCNU-induced oxidative stress and cell death. Apoptosis. 2013;18(11):1416-1425
  136. 136. Shi J, Sun B, Shi W, Zuo H, Cui D, Ni L, et al. Decreasing GSH and increasing ROS in chemosensitivity gliomas with IDH1 mutation. Tumour Biology. 2015;36(2):655-662
  137. 137. Ohba S, Mukherjee J, See WL, Pieper RO. Mutant IDH1-driven cellular transformation increases RAD51-mediated homologous recombination and temozolomide resistance. Cancer Research. 2014;74(17):4836-4844
  138. 138. Koncar RF, Chu Z, Romick-Rosendale LE, Wells SI, Chan TA, Qi X, et al. PLK1 inhibition enhances temozolomide efficacy in IDH1 mutant gliomas. Oncotarget. 2017;8(9):15827-15837
  139. 139. Trivedi RN, Almeida KH, Fornsaglio JL, Schamus S, Sobol RW. The role of base excision repair in the sensitivity and resistance to temozolomide-mediated cell death. Cancer Research. 2005;65(14):6394-6400
  140. 140. Yoshimoto K, Mizoguchi M, Hata N, Murata H, Hatae R, Amano T, et al. Complex DNA repair pathways as possible therapeutic targets to overcome temozolomide resistance in glioblastoma. Frontiers in Oncology. 2012;2:186
  141. 141. Tateishi K, Wakimoto H, Iafrate AJ, Tanaka S, Loebel F, Lelic N, et al. Extreme Vulnerability of IDH1 Mutant Cancers to NAD+ Depletion. Cancer Cell. 2015;28(6):773-784
  142. 142. Tateishi K, Higuchi F, Miller J, Koerner MVA, Lelic N, Shankar GM, et al. The alkylating chemotherapeutic temozolomide induces metabolic stress in IDH1-mutant cancers and potentiates NAD+ depletion-mediated cytotoxicity. Cancer Research. 2017;77(15):4102-4115
  143. 143. Lu Y, Kwintkiewicz J, Liu Y, Tech K, Frady LN, Su YT, et al. Chemosensitivity of IDH1 mutant gliomas due to an impairment in PARP1-mediated DNA repair. Cancer Research. 2017;77(7):1709-1718
  144. 144. Sulkowski PL, Corso CD, Robinson ND, Scanlon SE, Purshouse KR, Bai H, et al. 2-Hydroxyglutarate produced by neomorphic IDH mutations suppresses homologous recombination and induces PARP inhibitor sensitivity. Science Translational Medicine. 2017;9(375)
  145. 145. Inoue S, Li WY, Tseng A, Beerman I, Elia AJ, Bendall SC, et al. Mutant IDH1 downregulates ATM and alters DNA repair and sensitivity to DNA damage independent of TET2. Cancer Cell. 2016;30(2):337-348
  146. 146. Wang F, Travins J, DeLaBarre B, Penard-Lacronique V, Schalm S, Hansen E, et al. Targeted inhibition of mutant IDH2 in leukemia cells induces cellular differentiation. Science. 2013;340(6132):622-626
  147. 147. Kernytsky A, Wang F, Hansen E, Schalm S, Straley K, Gliser C, et al. IDH2 mutation-induced histone and DNA hypermethylation is progressively reversed by small-molecule inhibition. Blood. 2015;125(2):296-303
  148. 148. Okoye-Okafor UC, Bartholdy B, Cartier J, Gao EN, Pietrak B, Rendina AR, et al. New IDH1 mutant inhibitors for treatment of acute myeloid leukemia. Nature Chemical Biology. 2015;11(11):878-886
  149. 149. Kats LM, Reschke M, Taulli R, Pozdnyakova O, Burgess K, Bhargava P, et al. Proto-oncogenic role of mutant IDH2 in leukemia initiation and maintenance. Cell Stem Cell. 2014;14(3):329-341
  150. 150. DiNardo CD, Stein EM, de Botton S, Roboz GJ, Altman JK, Mims AS, et al. Durable remissions with ivosidenib in IDH1-mutated relapsed or refractory AML. The New England Journal of Medicine. 2018;378(25):2386-2398
  151. 151. Stein EM, DiNardo CD, Pollyea DA, Fathi AT, Roboz GJ, Altman JK, et al. Enasidenib in mutant IDH2 relapsed or refractory acute myeloid leukemia. Blood. 2017;130(6):722-731
  152. 152. Pusch S, Krausert S, Fischer V, Balss J, Ott M, Schrimpf D, et al. Pan-mutant IDH1 inhibitor BAY 1436032 for effective treatment of IDH1 mutant astrocytoma in vivo. Acta Neuropathologica. 2017;133(4):629-644
  153. 153. Kohanbash G, Carrera DA, Shrivastav S, Ahn BJ, Jahan N, Mazor T, et al. Isocitrate dehydrogenase mutations suppress STAT1 and CD8+ T cell accumulation in gliomas. The Journal of Clinical Investigation. 2017;127(4):1425-1437
  154. 154. Davis MI, Gross S, Shen M, Straley KS, Pragani R, Lea WA, et al. Biochemical, cellular, and biophysical characterization of a potent inhibitor of mutant isocitrate dehydrogenase IDH1. The Journal of Biological Chemistry. 2014;289(20):13717-13725
  155. 155. Suijker J, Oosting J, Koornneef A, Struys EA, Salomons GS, Schaap FG, et al. Inhibition of mutant IDH1 decreases D-2-HG levels without affecting tumorigenic properties of chondrosarcoma cell lines. Oncotarget. 2015;6(14):12505-12519
  156. 156. Turcan S, Makarov V, Taranda J, Wang Y, Fabius AWM, Wu W, et al. Mutant-IDH1-dependent chromatin state reprogramming, reversibility, and persistence. Nature Genetics. 2018;50(1):62-72
  157. 157. Johannessen TA, Mukherjee J, Viswanath P, Ohba S, Ronen SM, Bjerkvig R, et al. Rapid conversion of mutant IDH1 from driver to passenger in a model of human gliomagenesis. Molecular Cancer Research. 2016;14(10):976-983
  158. 158. Fujii T, Khawaja MR, DiNardo CD, Atkins JT, Janku F. Targeting isocitrate dehydrogenase (IDH) in cancer. Discovery Medicine. 2016;21(117):373-380
  159. 159. Borodovsky A, Salmasi V, Turcan S, Fabius AW, Baia GS, Eberhart CG, et al. 5-azacytidine reduces methylation, promotes differentiation and induces tumor regression in a patient-derived IDH1 mutant glioma xenograft. Oncotarget. 2013;4(10):1737-1747
  160. 160. Yamashita AS, da Costa Rosa M, Borodovsky A, Festuccia WT, Chan T, Riggins GJ. Demethylation and epigenetic modification with 5-azacytidine reduces IDH1 mutant glioma growth in combination with temozolomide. Neuro-Oncology. 2018
  161. 161. Chan SM, Thomas D, Corces-Zimmerman MR, Xavy S, Rastogi S, Hong WJ, et al. Isocitrate dehydrogenase 1 and 2 mutations induce BCL-2 dependence in acute myeloid leukemia. Nature Medicine. 2015;21(2):178-184
  162. 162. Karpel-Massler G, Ishida CT, Bianchetti E, Zhang Y, Shu C, Tsujiuchi T, et al. Induction of synthetic lethality in IDH1-mutated gliomas through inhibition of Bcl-xL. Nature Communications. 2017;8(1):1067
  163. 163. Flynn RL, Cox KE, Jeitany M, Wakimoto H, Bryll AR, Ganem NJ, et al. Alternative lengthening of telomeres renders cancer cells hypersensitive to ATR inhibitors. Science. 2015;347(6219):273-277
  164. 164. Spino M, Kurz SC, Chiriboga L, Serrano J, Zeck B, Sen N, et al. Cell surface Notch ligand DLL3 is a therapeutic target in isocitrate dehydrogenase mutant glioma. Clinical Cancer Research. 2019;25(4):1261-1271
  165. 165. Schumacher T, Bunse L, Pusch S, Sahm F, Wiestler B, Quandt J, et al. A vaccine targeting mutant IDH1 induces antitumour immunity. Nature. 2014;512(7514):324-327
  166. 166. Pellegatta S, Valletta L, Corbetta C, Patane M, Zucca I, Riccardi Sirtori F, et al. Effective immuno-targeting of the IDH1 mutation R132H in a murine model of intracranial glioma. Acta Neuropathologica Communications. 2015;3:4
  167. 167. Amankulor NM, Kim Y, Arora S, Kargl J, Szulzewsky F, Hanke M, et al. Mutant IDH1 regulates the tumor-associated immune system in gliomas. Genes & Development. 2017;31(8):774-786
  168. 168. Berghoff AS, Kiesel B, Widhalm G, Wilhelm D, Rajky O, Kurscheid S, et al. Correlation of immune phenotype with IDH mutation in diffuse glioma. Neuro-Oncology. 2017;19(11):1460-1468
  169. 169. Seltzer MJ, Bennett BD, Joshi AD, Gao P, Thomas AG, Ferraris DV, et al. Inhibition of glutaminase preferentially slows growth of glioma cells with mutant IDH1. Cancer Research. 2010;70(22):8981-8987
  170. 170. Emadi A, Jun SA, Tsukamoto T, Fathi AT, Minden MD, Dang CV. Inhibition of glutaminase selectively suppresses the growth of primary acute myeloid leukemia cells with IDH mutations. Experimental Hematology. 2014;42(4):247-251
  171. 171. Cuyas E, Fernandez-Arroyo S, Corominas-Faja B, Rodriguez-Gallego E, Bosch-Barrera J, Martin-Castillo B, et al. Oncometabolic mutation IDH1 R132H confers a metformin-hypersensitive phenotype. Oncotarget. 2015;6(14):12279-12296
  172. 172. Odia Y, Orr BA, Bell WR, Eberhart CG, Rodriguez FJ. cMYC expression in infiltrating gliomas: Associations with IDH1 mutations, clinicopathologic features and outcome. Journal of Neuro-Oncology. 2013;115(2):249-259
  173. 173. Chipumuro E, Marco E, Christensen CL, Kwiatkowski N, Zhang T, Hatheway CM, et al. CDK7 inhibition suppresses super-enhancer-linked oncogenic transcription in MYCN-driven cancer. Cell. 2014;159(5):1126-1139
  174. 174. Christensen CL, Kwiatkowski N, Abraham BJ, Carretero J, Al-Shahrour F, Zhang T, et al. Targeting transcriptional addictions in small cell lung cancer with a covalent CDK7 inhibitor. Cancer Cell. 2014;26(6):909-922
  175. 175. Tateishi K, Iafrate AJ, Ho Q , Curry WT, Batchelor TT, Flaherty KT, et al. Myc-driven glycolysis is a therapeutic target in glioblastoma. Clinical Cancer Research. 2016;22(17):4452-4465

Written By

Kensuke Tateishi and Tetsuya Yamamoto

Submitted: 08 October 2018 Reviewed: 18 January 2019 Published: 23 February 2019