Next Article in Journal
Identification and New Indication of Melanin-Concentrating Hormone Receptor 1 (MCHR1) Antagonist Derived from Machine Learning and Transcriptome-Based Drug Repositioning Approaches
Previous Article in Journal
Foliar Silicon Spray before Summer Cutting Propagation Enhances Resistance to Powdery Mildew of Daughter Plants
Previous Article in Special Issue
Biosynthesis and Degradation of Sulfur Modifications in tRNAs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

FTO m6A Demethylase in Obesity and Cancer: Implications and Underlying Molecular Mechanisms

by
Sarah Kassem Azzam
1,2,3,
Habiba Alsafar
1,3,4,5 and
Abdulrahim A. Sajini
1,2,*
1
Department of Biomedical Engineering, Khalifa University of Science and Technology, Abu Dhabi P.O. Box 127788, United Arab Emirates
2
Healthcare Engineering Innovation Center (HEIC), Department of Biomedical Engineering, Khalifa University, Abu Dhabi P.O. Box 127788, United Arab Emirates
3
Center for Biotechnology, Khalifa University of Science and Technology, Abu Dhabi P.O. Box 127788, United Arab Emirates
4
Department of Genetics and Molecular Biology, College of Medicine and Health Sciences, Khalifa University of Science and Technology, Abu Dhabi P.O. Box 127788, United Arab Emirates
5
Emirates Bio-Research Center, Ministry of Interior, Abu Dhabi P.O. Box 389, United Arab Emirates
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(7), 3800; https://doi.org/10.3390/ijms23073800
Submission received: 29 January 2022 / Revised: 20 March 2022 / Accepted: 25 March 2022 / Published: 30 March 2022

Abstract

:
Fat mass and obesity-associated protein (FTO) is the first reported RNA N6-methyladenosine (m6A) demethylase in eukaryotic cells. m6A is considered as the most abundant mRNA internal modification, which modulates several cellular processes including alternative splicing, stability, and expression. Genome-wide association studies (GWAS) identified single-nucleotide polymorphisms (SNPs) within FTO to be associated with obesity, as well as cancer including endometrial cancer, breast cancer, pancreatic cancer, and melanoma. Since the initial classification of FTO as an m6A demethylase, various studies started to unravel a connection between FTO’s demethylase activity and the susceptibility to obesity on the molecular level. FTO was found to facilitate adipogenesis, by regulating adipogenic pathways and inducing pre-adipocyte differentiation. FTO has also been investigated in tumorigenesis, where emerging studies suggest m6A and FTO levels are dysregulated in various cancers, including acute myeloid leukemia (AML), glioblastoma, cervical squamous cell carcinoma (CSCC), breast cancer, and melanoma. Here we review the molecular bases of m6A in tumorigenesis and adipogenesis while highlighting the controversial role of FTO in obesity. We provide recent findings confirming FTO’s causative link to obesity and discuss novel approaches using RNA demethylase inhibitors as targeted oncotherapies. Our review aims to confirm m6A demethylation as a risk factor in obesity and provoke new research in FTO and human disorders.

1. Introduction

The fat mass and obesity-associated (FTO) gene was the first gene to show the strongest association with polygenic obesity. The FTO protein encoded by the FTO gene belongs to a family of dioxygenases that depend on Fe (II) and 2-oxoglutarate (2OG) for their enzymatic activity [1]. Enzymatically, FTO is an RNA N6-methyladenosine (m6A) demethylase in eukaryotic cells [2]. m6A is the most abundant internal post-transcriptional modification in messenger RNA (mRNA) with high conservation across human and mouse [3]. FTO was later identified to also demethylate N6,2′-O-dimethyladenosine (m6Am)-dimethyl-modified mRNAs and reduce their chemical stability [4]. m6A is a highly-dynamic and reversible modification with writers depositing them on RNA while erasers remove them [5,6]. Once m6A is deposited on RNA, different groups of proteins, called readers, facilitate downstream functions. Several cellular processes are modulated via m6A including alternative splicing, mRNA stability, translation, transportation, degradation, and gene regulation [3,7,8,9,10,11]. Additionally, the co-localization of FTO and components of m6A methylation complex with splicing proteins in nuclear speckles, which are involved in the assembly of mRNA-processing factors, further supports regulatory roles of m6A in RNA processing [2,12]. The imbalance of m6A has been associated with various human disorders, such as obesity, diabetes, heart failure, brain diseases, and various cancers [13,14,15,16,17,18]. In GWAS, FTO constantly scored high among obese and cancer cohorts. Molecularly, variation in FTO could explain the m6A perturbation found in both obesity and cancer. Hence, many studies have explored the molecular mechanisms of FTO in adipogenesis and tumorigenesis with conflicting results. The inconsistency arises from the fact that most FTO single-nucleotide polymorphisms (SNPs) associated with overweight are intronic. This has divided scientists studying FTO in obesity into groups hypothesizing that FTO introns act as cis-regulatory sites for adjacent genes, while other groups hypothesize that FTO introns act as auto-regulators. Here we review studies from both groups and support the fact that FTO does directly modulate obesity on the m6A level. We also discuss FTO’s mechanistic role in cancer and shed some light on m6A demethylase inhibitors as novel cancer-targeted therapies. We hope our review encourages aligned views in studies involving FTO, obesity, and cancer.

2. FTO Association with Obesity as an m6A Demethylase (Obesogenic Role)

2.1. FTO Association with Obesity in Genome-Wide Association Studies and Animal Models

FTO was the first gene having the strongest genetic association with polygenic obesity. Using GWAS, scientists identified significant SNPs in the first intron of FTO that were associated with obesity. These initial screens were done in 2007 among European populations [19,20,21], and later confirmed in other populations, including Hispanics [22,23], East Asians [24,25,26], Native Americans [27], South Asians [28,29], Africans [30] and Middle-Easterns [31,32,33,34,35,36]. Genomically, intron 1 and 2 of the FTO gene show the strongest genome- association with obesity, where 89 variants were identified [19,21]. Phenotypically, the higher body-mass index (BMI) recorded in individuals carrying FTO risk variants may be attributed to increased energy intake [37,38,39] and reduced food satiety [40], yet not due to impaired energy expenditure [37,41]. Given the obstacles of exploring FTO in humans, researchers used mouse models to delineate FTO’s roles in energy homeostasis and fat-mass regulation [13,42,43,44,45,46,47]. In mice, lower fat mass and body weight were reported when FTO was knocked out or ablated by loss-of-function mutations [42,43,44,45]. Conversely, FTO overexpression led to higher body and fat mass, as seen from knock-in mice models as well as plants’ increased yield and biomass, upon FTO overexpression [13,48]. Apart from weight, mice lacking FTO also suffered from postnatal growth retardation, which highlights FTO’s role in the central nervous system development [42,44]. In pigs, FTO knockdowns reduced lipid accumulation supporting FTO’s role in adipogenesis [49]. Humans homozygous for FTO mutations that deactivate it enzymatically, suffer from multiple congenital abnormalities, severe growth retardation and die in infancy, highlighting an essential role of FTO in normal human development [50,51]. Interestingly, none of the heterozygous parents of the aforementioned FTO-enzymatic mutant were clinically obese, suggesting a single functional copy of FTO was sufficient to minimize the risk of obesity [50]. However, further studies investigating the relation between FTO heterozygosity and obesity are needed. All in all, data across different species support FTO’s involvement in higher body mass.

2.2. FTO Molecular Association with Obesity as an m6A Demethylase

GWAS represent a powerful tool in establishing connections between FTO variants and phenotypical features of obesity such as BMI and waist circumference. Nevertheless, translating the findings from such studies to molecular roles of FTO variants in obesity is a continuous challenge. Since the initial identification of FTO as an m6A demethylase, a number of recent studies began to unravel a connection between FTO-dependent m6A demethylation and obesity. Karra et al. uncovered a novel role of FTO as a potential regulator of ghrelin, the hunger hormone, by comparing humans homozygous for FTO SNP rs9939609 risk allele, with wild-type carriers [52]. They found that blood cells carrying the risk allele expressed elevated FTO mRNA levels which in turn led to higher ghrelin expression [52]. This study highlighted a potential mechanism underlying the obesity-prone eating behavior observed in FTO rs9939609 individuals. However, further experiments are required to confirm whether modulating ghrelin is sufficient to explain FTO’s molecular role in obesity.
Using porcine adipocytes, m6A levels were shown to negatively regulate lipid accumulation [49]. The inverse regulatory effect of m6A was confirmed via FTO overexpression and knockdown studies, as well as by exogenously regulating m6A levels using cycloleucine, a chemical inhibitor of RNA methylation [49,53]. A study by Zhao et al. [10] using mice 3T3-L1 pre-adipocytes reported an inverse correlation during adipogenesis between m6A levels and FTO expression. Specifically, m6A demethylation by FTO weakens the binding affinity of the splicing regulatory (SR) protein SRSF2. The Runt-related transcription factor 1 (RUNX1T1), an adipogenic gene, is a target of SRSF2, and has two splice variants; a short (S) one and a long (L) one.
The perturbed binding of SRSF2 results in the skipping of exon 6 and the generation of the RUNX1T1 S variant, which induces pre-adipocyte differentiation. In contrast, FTO knockdown leads to m6A-accumulation and recruitment of SRSF2 towards RUNX1T1 transcripts. SRSF2 assists in generating the RUNX1T1 L variant by including exon 6 [10]. This example illustrates how FTO demethylation of m6A regulates RNA processing and pre-adipocyte differentiation via alternative splicing. Another study by Zhang et al. [54] also highlighted the role of FTO demethylation in adipocytes differentiation by using murine 3T3-L1 pre-adipocytes. They showed that FTO knockdown perturbed 3T3-L1 pre-adipocytes differentiation into mature adipocytes. Conversely, FTO overexpression in 3T3-L1 pre-adipocytes increased adipogenesis by lowering m6A levels. Interestingly, the expression of an FTO missense mutant, lacking its catalytic activity (R96Q), impeded differentiation but had no effect on m6A levels; potentially explained via competing with endogenous FTO for substrate binding [54]. The mutation of Arg96 to Gln in the R96Q FTO mutant compromises its demethylase activity; given the essential role of Arg96 as an interacting residue in substrate binding [55]. The use of Rosiglitazone, a PPARγ agonist, in R96Q mutant expressing cells was shown to partially overcome adipogenesis inhibition; suggesting that FTO plays a role in adipogenesis at least partially via PPARγ, a master adipogenic regulator [54].
In-vitro studies connecting FTO to adipogenesis were further confirmed by Merkestein et al. [46]. Their study provided mechanistic insights relating FTO upregulation with obesity; where it regulates early adipogenesis during the mitotic clonal expansion (MCE) phase. In line with a previous study [10], FTO is reported to regulate the splicing of the adipogenesis-related transcription factor RUNX1T1. Elevated pro-adipogenic RUNX1T1 S-isoform stimulates the MCE phase in adipocytes. Stimulating the MCE phase, through FTO, enhances adipogenic differentiation via increasing D-type Cyclin genes (Cyclin D1 and Cyclin D3) and PPARγ and C/EBPα, master regulators of adipogenesis. Merkestein et al. showed that FTO knockdown after MCE (48 h post adipogenic induction) did not influence adipogenic genes; highlighting the time-sensitive role of FTO in adipogenesis. In-vivo studies of mice linked the role of high-fat diet (HFD) in FTO-overexpressing mice (FTO-4), as fat pads from such mice displayed increased adipocytes number, indicating hyperplasia induction upon HFD [46].
Recent studies showed conflicting results in regards to the effect of FTO SNPs on FTO expression and obesity development. Some studies indicate that FTO SNPs work by modulating adjacent genes such as IRX3 and RPGRIP1L [56,57]. Other studies show that FTO SNPs increased FTO expression, hence act as autoregulators [52,58]. The study by Merkestein et al. supported the role of FTO in adipogenesis independent of either IRX3 or RPGRIP1L [46]; implying that gain-of-function variants in FTO may stimulate adipogenesis in carriers of FTO variant on an HFD.
The involvement of FTO in energy expenditure regulation was pointed on a molecular level, via the browning effect of white adipose tissue in mice and humans [59,60]. Tews et al. demonstrated that FTO deficiency results in the upregulation of uncoupling protein 1 (UCP-1) in adipocytes; hence, enhancing mitochondrial uncoupling and energy expenditure in brown adipocytes [59]. This raises the speculation that FTO may increase body weight via reducing UCP-1 expression and inhibiting the browning of white adipose tissue. Claussnitzer group [60] studied the mechanism of FTO variant (T-to-C) rs1421085, in connection with obesity; as it was shown to disrupt a conserved repressor motif (ARID5B). The loss of the later motif causes an upregulation of downstream targets IRX3 and IRX5 during early adipocyte differentiation. Consequently, a reduction in heat dissipation and loss of mitochondrial thermogenesis occurs (a process which is regulated via UCP1, peroxisome proliferator-activated receptor γ coactivator 1 α (PGC1α), and PR-domain containing 16 (PRDM16)). This shift from energy-dissipating beige adipocytes into energy-storing white adipocytes results in an increased lipid storage capacity and weight gain [60]. Studies linking FTO expression with white adipose tissue browning represent novel avenues for better obesity management in the future [59,60].
FTO activity as an m6A demethylase is also correlated with the regulation of lipid accumulation in skeletal muscle cells via its regulatory role in the AMP-activated protein kinase (AMPK) signaling pathway [61]. The energy sensor AMPK negatively regulates lipid accumulation in skeletal muscle cells, at least partially via regulating FTO expression levels and FTO-dependent m6A demethylation. Active AMPK mitigates FTO expression, resulting in higher m6A methylation levels (positive regulation) in skeletal muscles which reduces lipid accumulation (negative regulation). Wu and colleagues [61] showed that m6A modifications increase lipid accumulation via decreasing lipolysis and oxidation-associated transcripts (adipose triglyceride lipase (ATGL), hormone-sensitive lipase (HSL), and PGC1 α) while stabilizing lipid synthase related transcripts (CCAAT/enhancer binding protein α (C/EBPα) and fatty acid synthase (FAS)). These regulatory mechanisms could act as manipulation tools for lipid deposition in skeletal muscle by targeting AMPK and/or utilizing m6A-associated drugs [61].
Moreover, FTO seems to play a regulatory role in cell cycle progression by modulating m6A and its reader YTHDF2 as reported by Wu et al. [62]. FTO was initially reported to regulate early adipogenesis by regulating MCE, a stage occurring within 48 h of adipogenic induction [46]. In 2018, Wu et al. provided further mechanistic details, whereby FTO knockdown resulted in higher m6A levels on cyclin A2 mRNA (CCNA2) and cyclin-dependent kinase 2 mRNA (CDK2). CCNA2 and CDK2 are two key regulators of early mitotic events, where their cell cycle role is focused on mediating the S to G2 phase transition [63]. The accumulation of m6A on both CCNA2 and CDK2 mRNAs signals degradation via YTHDF2; thus impairing cell cycle progression and suppressing adipogenesis in 3T3-L1 pre-adipocytes [62]. As an m6A reader, YTHDF2 was previously reported to regulate mRNA degradation [7]. It would be interesting to explore the roles of other m6A- readers in adipogenesis such as YTHDF1 and YTHDF3, given their functions in translation [8,64].
Recently, the Zinc finger protein 217 (Zfp217) was uncovered as another player in FTO-dependent adipogenic regulation [65]. Zfp217 is an oncogenic protein overexpressed in multiple human cancers [66,67,68]. Transcriptionally, Zfp217 binds to FTO’s promoter region and increases its expression; thereby reducing global m6A levels. Zfp217 ablation in 3T3-L1 pre-adipocytes halts adipogenic differentiation by decreasing the expression of key adipogenic genes namely PPARγ and aP2. In addition to its transcriptional activation role of FTO, Zfp217 interacts with YTHDF2 to modulate mRNA m6A modifications. Specifically, Zfp217 sequesters YTHDF2 and shifts the equilibrium in YTHDF2-FTO competition over m6A sites, by favoring FTO m6A demethylation activity. This study by Song et al. supports previous findings linking Zfp217 overexpression with adipogenesis; whereby Zfp217 promotes adipogenesis via transcriptional and post-transcriptional mechanisms [65,69].
Another FTO-dependent factor in adipogenesis is nicotinamide adenine dinucleotide phosphate (NADP); highlighting a direct association between metabolism and RNA m6A demethylation [70]. In 3T3-L1 pre-adipocytes NADP directly binds to FTO and enhances its demethylase activity; hence promoting m6A removal and adipogenesis. In this study, both control and FTO knockdown 3T3-L1 pre-adipocytes were treated with NADPH prior to adipogenic induction; and notably, NADPH enhanced adipogenesis in control pre-adipocytes. However, FTO knockdown impaired NADPH’s ability to enhance adipogenesis, as shown from lipid staining and triglyceride accumulation. In line with 3T3-L1 pre-adipocytes results, Wang et al. also demonstrated that NADP modulated m6A levels in FTO-knockout mice [70]. It will be interesting to explore how NADP enhances FTO and whether this process is found in human adipocytes.
Novel mechanistic insights were recently discovered by Wang et al. [71] which connects m6A modifications with adipogenesis and autophagy. Specifically, FTO knockdown in 3T3-L1 pre-adipocytes and in porcine primary pre-adipocytes resulted in higher m6A levels on two autophagy-related genes ATG5 and ATG7; hence marking them for degradation by YTHDF2. Low ATG5 and ATG7 proteins lead to autophagy and adipogenesis inhibition. Pinpointing ATG5 and ATG7 as key regulators in FTO-dependent autophagy and adipogenesis further contributes to our understanding of how FTO could be regulating obesity. Recently, Wu et al. unveiled how FTO can promote thermogenesis and white-to-beige adipocyte transition. Mechanistically, FTO loss in adipose tissue increases m6A accumulation on hypoxia inducible factor 1 subunit alpha (HIF1A) mRNAs. Abundant m6A enhances HIF1A translation by m6A-reader YTHDC2. Transcriptionally, HIF1A activates thermogenic genes such as PGC1A, PRDM16, and PPARG, thereby promoting a “browning” process of white adipose tissue [72].
In Table 1 below, we summarize a list of studies classifying FTO as an obesogenic regulator and provide an illustrative summary of those studies in Figure 1.

3. FTO m6A Demethylase Association with Various Cancers

3.1. FTO Is Overexpressed in a Number of Human Cancers

Further to its association with obesity, epidemiological studies link FTO SNPs with higher risks of endometrial cancer [74,75,76], breast cancer [77,78], pancreatic cancer [76,79,80] and melanoma [81]; with some variations among different races. Some of FTO SNPs were reported to be BMI-related [75,77,79] while others were independent of BMI [74,78,80,81]. These contradicting results contribute further to the complexity of the re-lationship between obesity and cancer. Emerging studies observe that m6A modifications and their modifying enzymes are dysregulated in various cancers, specifically acute myeloid leukemia (AML), glioblastoma, cervical squamous cell carcinoma (CSCC), breast cancer and melanoma. These observations support FTO and m6A modifications in tu-morigenesis, hence making FTO a potential target for precision medicine in cancer, as shall be discuss herein.

3.2. Preliminary Mechanistic Investigations of FTO Oncogenic Role

The connection between endometrial cancer and FTO variants [74,75] were first reported in 2011 by Zhang et al. [82]. In particular, FTO expression was constantly abundant in both endometrial cancer specimens and endometrial cancer cell lines. It was later discovered that β-estradiol (E2) drove FTO overexpression by binding to estrogen receptor (ER), which then modulated FTO by activating PI3K/AKT and MAPK pathways. This in turn upregulates cyclin D1 and metalloproteinases MMP2/9 [82]. Interestingly, the association between FTO and endometrial carcinoma did not consider FTO as an m6A demethylase, since FTOs demethylation activity was not reported until later that same year [2]. In breast cancer FTO is also overexpressed in mastectomy tissue; specifically in HER-2 high subtypes [83]. Clinically, FTO abundance is associated with poor prognosis and early relapse of endometrial cancer [84]. Additionally, estrogen activates mTOR signaling [84], a pathway implicated in endometrial cancer development and progression [85]. Thus, leading to FTO nuclear localization, and increasing endometrial cancer proliferation [84].
The oncogenic role of FTO in breast cancer was studied by comparing breast cancer cell lines to normal breast cells. This comparison revealed that FTO was overexpressed in breast cancer cells; denoting it as a potential diagnostic marker in breast cancer. Mechanistically, FTO overexpression in breast cancer cells increases energy metabolism via the PI3K/AKT signaling pathway, which is abnormally activated in many human cancers [86,87,88]. FTO overexpression upregulates pyruvate kinase and hexokinase, and increases ATP generation; thus, promoting aerobic glycolysis [86]. The FTO oncogenic role was also reported in human gastric cancer (GC), where FTO was upregulated on both protein and mRNA levels. Similar to breast cancer FTO abundance is associated in GC with metastasis; marking it as a poor prognostic marker [89].

3.3. The Oncogenic Role of FTO as an m6A Demethylase and Associations with Cancers

The mechanistic roles of FTO in tumorigenesis are currently being elucidated by various research groups in terms of its m6A demethylase activity.
A study by Li et al. [90] discovered a previously unrecognized oncogenic role of FTO in acute myeloid leukemia (AML). They showed that FTO was abnormally regulated in subtypes of AML, specifically t(11q23)/MLL-rearranged, t(15;17)/PML-RARA, FLT3-ITD, and/or NPM1-mutated AMLs. FTO expression enhanced AML cellular viability and proliferation while inhibiting apoptosis as well as significantly promoting Leukemogenesis in-vivo [90]. Furthermore, depletion of FTO significantly enhanced all-trans-retinoic acid (ATRA)-induced AML cell differentiation. Here, FTO regulates the mRNA stability of two genes: Ankyrin repeat and SOCS box protein 2 (ASB2) and retinoic acid receptor α (RARA) by controlling m6A levels. Interestingly, it was demonstrated that FTO’s oncogenic role was independent of YTHDF2 and YTHDF1; indicating that an alternative m6A reading process might be involved [90]. Cui et al. established a causative link between m6A modifications and glioblastoma, a highly aggressive form of brain cancer [16]. They show that the use of FTO inhibitor (Meclofenamic acid MA) [91] to target glioblastoma stem cells (GSCs), results in suppressing GSC-induced tumorigenesis as well as increasing the lifespan of GSC-engrafted mice [16]. These findings suggested that m6A erasers, like FTO, represent interesting targets for glioblastoma therapy. Su et al. [92] report the use of R-2-hydroxyglutarate (R-2HG), a previously reported oncometabolite [93,94], as a small-molecule inhibitor of FTO; displaying anti-proliferative effects in leukemia and glioma. In fact, R-2HG exerts its anti-tumor effects via inhibiting FTO enzymatic activity [92]. This in turn increases m6A levels on FTO oncogenic mRNA targets, specifically reported in this study MYC/CEBPA; hence reducing their stability and suppressing relevant oncogenic pathways. Moreover, Su et al. reported that R-2HG administration resulted in synergistic interactions and displayed better therapeutic effects when combined with first-line chemotherapy drugs such as daunorubicin or decitabine, as verified in-vivo using leukemia mice models xeno-transplanted with R-2HG sensitive leukemic cells [92].
Zhou et al. study [95] was the first to report FTO’s role in treatment response of a solid tumor, specifically cervical squamous cell carcinoma (CSCC). In their study, they suggest that FTO/β-catenin/ERCC1 axis plays a critical role in regulating chemo-radiotherapy resistance in CSCC. FTO mRNA levels were upregulated in CSCC tissues relative to adjacent normal tissues. FTO promoted chemo-radiotherapy resistance in-vitro and in-vivo, via demethylating β-catenin transcripts; thus upregulating its mRNA and protein levels. Subsequently, excision repair cross-complementation group 1 (ERCC1) was found to be an activated downstream target contributing to the FTO/β-catenin induced chemo-radiotherapy resistance in CSCC cells (Figure 2) [95]. FTO serves as an oncogenic regulator, by modulating m6A levels on MYC and E2F1; thus increasing cervical cancer’s migration and proliferation [96].
FTO has a pro-tumorigenic role in lung cancer, as evident from its overexpression in lung cancer tissues and cell lines [97,98]. FTO wildtype overexpression, but not its mutant enzymatically-inactive forms, promoted oncogenic functions in lung cancer cells, including proliferation, invasion, and colony formation ability. Mechanistic analyses revealed two functional targets of FTO; first, upregulation of Myeloid Zinc Finger Protein 1 (MZF1) in lung squamous cell carcinoma (LUSC), and second: ubiquitin-specific protease-7 (USP7) in non-small cell lung cancer (NSCLC). Both targets’ stability was dependent on FTO m6A demethylase activity. FTO upregulation in lung cancer cells and the resulting reduced m6A levels on MZF1 and USP7 increased their stability; thereby, promoting oncogenic functions [97,98].In colon cancer cell lines, Tsuruta et al. reported the upregulation of FTO and its target: Programmed death 1 ligand (PD-L1) [99]. PD-L1 is one of the immune-check point molecules, which when highly expressed on tumor cells permits their escape from immune surveillance [100].
FTO was also upregulated in human breast cancer [101]. Niu et al. demonstrated that FTO promoted breast cancer cell proliferation, colony formation, and metastasis. Mechanistically, FTO exerted its oncogenic role in breast cancer via demethylating BCL2 Interacting Protein 3 (BNIP3) [101], a pro-apoptosis member in the Bcl-2 apoptotic protein family [102]. Consequently, downregulating BNIP3 and resulting in reduced apoptosis of breast cancer cells. FTO demethylase action on BNIP3 negatively regulated its expression and modulated its stability, through a YTHDF2-independent manner [101]. This is in line with an earlier study which reported the regulation of AML associated oncogenic target genes ASB2 and RARA by FTO via a YTHDF2-independent mechanism [90]. This indicates that an alternative reading process of m6A-modified transcripts and thus alternative stabilizing exists and ought to be explored.
A pro-tumorigenic role of FTO was recently discussed in melanoma, illustrating the upregulated FTO levels in human melanoma samples and multiple melanoma cell lines, and reporting that FTO promotes melanoma tumor growth in-vivo [103]. Molecular details were discussed by Yang et al. [103], in which FTO knockdown resulted in m6A hyper-methylated melanoma-promoting genes namely, PD-1 (PDCD1), CXCR4, and SOX10; consequently reducing their mRNA stability, at least partially via a YTHDF2-mediated mechanism. The reported cascade of events relating FTO upregulation to melanoma tumorigenesis is suggested to be initiated by metabolic stress, a metabolic challenge faced by tumor cells, which in turn induces FTO upregulation as a means of melanoma adaptation to metabolic challenges. Furthermore, FTO knockdown alleviated melanoma resistance to anti-PD-1 treatment; suggesting that FTO inhibition in combination with anti-PD-1 immunotherapy is a promising anti-melanoma therapy [103].
In line with the reported oncogenic role of FTO demethylase in a subset of AMLs [90] and the anti-tumor effect exerted by an FTO inhibitor (R-2HG) [92], Huang et al. [104] utilized structure-guided design to optimize Meclofenamic acid (MA), a previously identified FTO inhibitor [91]. This led to developing FB23 and FB23-2, as two more potent inhibitors of FTO-mediated demethylation [104]. The MA-derived FB23-2 FTO inhibitor was shown to inhibit proliferation and to induce differentiation of human AML cell lines, as well as to inhibit AML progression in-vivo using mice models. The inhibitory effects were mediated by upregulating RARA and ASB2 (negative regulation) and downregulating MYC and CEPA (positive regulation), as a result of elevated m6A levels on transcripts following FTO inhibition (Figure 2) [104].
In endometrial carcinoma (EC), FTO overexpression promoted endometrial cancer proliferation and invasion, through PI3K/AKT and MAPK signaling pathways [82]. A recent study further pointed out that FTO overexpression in EC leads to m6A hypo-methylated HOXB13, a homeobox transcription factor. Hypo-methylated HOXB13 transcripts are protected from YTHDF2-mediated decay, thus promoting higher HOXB13 expression and activating the Wnt signaling pathway. This in turn promotes EC invasion and metastasis [105].
Alterations to m6A modification play various roles in cancers, as the methylated target genes can be proto-oncogenes or tumor suppresser genes, as presented with FTO differential expression levels in pancreatic cancer [106,107]. Tang et al. [107] reported FTO overexpression in pancreatic cancer. They further indicated that FTO knockdown led to a compromised proliferation in pancreatic cancer cells, and in terms of its m6A demethylase activity, led to reduced c-Myc mRNA transcript levels by modulating its m6A enrichment [107]. In this study, c-Myc oncogene was investigated for its reported regulatory role in the cell cycle by mediating entry into S-phase [108]. While all studies discussed thus far highlight FTO’s oncogenic role in various cancers, FTO’s tumor suppressor function was recently uncovered by Zeng et al. They report reduced FTO expression in pancreatic cancers, which increased m6A levels on PJA2 promoting its decay via YTHDF2. Suppression of PJA2 expression enhances Wnt signaling; thus promoting proliferation and invasion in pancreatic cancer cells [106]. These two studies on FTO’s role in pancreatic cancer cells present a dual role of FTO in pancreatic tumorigenesis that ought to be addressed, as it is reported as a tumor-promotor by one study [107] while its tumor-suppressor role is addressed in another one [106].
Interestingly, FTO’s tumor suppressor function was also reported in Ovarian Cancer Stem Cells (OCSCs) [109]. Huang et al. pointed to FTO downregulation in ovarian cancer cells and stem cells. Through transcriptome-wide RNA-seq and m6A mapping, two phosphodiesterase genes (PDE4B and PDE1C) with m6A hyper-methylated levels were identified. Subsequently, these stabilized m6A hyper-methylated transcripts promote the hydrolysis of the second messenger cAMP. Thus, downregulating cAMP and promoting tumorigenesis and self-renewal [109]. Interestingly, Huang et al. proposed the potential identification of hyper-methylated transcripts in ovarian cancer cells via IGF2BP1-3 as an RNA binding protein, which leads to increased transcript stability and gene expression levels [110] Overexpression of wild-type FTO was shown to otherwise inhibit ovarian cancer stemness and tumorigenesis, while mutant enzymatically-inactive forms failed to do so; highlighting the necessary role of FTO as m6A demethylase in ovarian cancer [109].
Overall, studies support FTO critical oncogenic roles in various cancer types as an m6A demethylase, given its abnormal expression and effect on many target genes. Figure 2 below presents a visual summary of studies on FTO demethylase activity, its overexpression in several human cancers, and the affected signaling pathways.

4. FTO Inhibition in Tackling FTO-Associated Disorders

The emergence of numerous GWA studies relating FTO SNPs with obesity/overweight across different populations motivated further research into investigating FTO physiological and pathological relevance; yet, the underlying molecular role of FTO in adipogenesis and cancer requires further research. A number of studies have reported the use of biochemical inhibitors of FTO, considering its pathological implications in obesity and cancers for instance.
Chen and team [111] identified the first potent inhibitor of FTO in 2012, the natural product Rhein. Rhein acts as a competitive inhibitor versus the nucleic acid substrate; thus, hindering FTO m6A demethylation activity. Rhein inhibitory activity of FTO was further affirmed on mRNA level using BE(2)-C cells (a clone of the SK-N-BE(2) neuroblastoma cell line), and illustrated significantly higher m6A levels [111]. Another study identified small molecules termed 2OG analogs as potent FTO inhibitors [112]. Their investigations were based on the fact that FTO demethylation activity is 2-oxoglutarate (2OG)-dependent [1]. In their study, they highlight the four most potent inhibitors, of which inhibitor 19 was previously reported in [111]. This study illustrated FTO demethylase activity using non-cellular in-vitro tests, whereby 3-methylthymidine (3meT) served as a test substrate. Further testing of FTO demethylating activity using other FTO substrates, such as m6A being the most abundant mRNA modification and the best substrate for FTO discovered yet under physiological conditions [2], would be required.
Another research study by Huang’s group [91] identified Meclofenamic acid (MA) as a selective inhibitor of FTO m6A demethylase activity over ALKBH5. MA as a nonsteroidal anti-inflammatory drug (NSAID) [113], was shown to inhibit FTO independent of 2OG or iron ions chelation. In-vitro tests revealed that MA increased m6A levels in HeLa cells mRNA via selectively inhibiting FTO, as affirmed from FTO overexpression and subsequent treatment with MA. Interestingly, the FTO-MA complex revealed a novel specific interaction called ‘a β-hairpin motif within nucleotide recognition lid (NRL)’, which explains the selective inhibition of FTO by MA [91].
Again, pertaining to the 2OG-dependent enzymatic activity of FTO demethylase [1], a new compound (molecule 7d) was developed by Zheng et al. [114] as an inhibitor of FTO enzymatic activity; inducing an increase in m6A total cellular RNA from HeLa cells. Their study presented 7d as the first FTO inhibitor with anticonvulsant activity, as affirmed from an animal model of pharmaco-resistant epilepsy (termed six Hz model) [114]; yet future investigations on this inhibitor effect on FTO activity in relation to obesity and cancers as FTO-associated disorders would be interesting. Toh and team [115] identified a selective inhibitor of FTO demethylase activity that illustrated increased m6A using a newly-synthesized compound 12. Interestingly, Glu234FTOresidue was found to be a key determinant for FTO affinity and specificity towards its substrates; as other equivalent residues from AlkB subfamilies had weakened or abolished interaction with compound 12 [115].
Another newly synthesized compound N-(5-Chloro-2,4-dihydroxyphenyl)-1 phenylcyclobutane-carboxamide (abbreviated N-CDPCB, 1a) was reported [116], whose structure bound to FTO partially overlaps with MA [91]. He and the team affirmed N-CDPCB inhibitory activity of FTO catalytic action, which showed that m6A levels were increased in mRNA of FTO-overexpressing 3T3-L1 pre-adipocytes as compared to untreated cells. They also reported preliminary inhibitory activity of FTO demethylase activity on 3meT as an initial testing substrate using non-cellular assays [116]. A recent study re-visited the use of Rhein as a putative inhibitor of FTO [117]. In their investigation, Huang et al. provided mechanistic insights into the cellular actions by which Rhein modulates adipogenesis. Interestingly, by comparing Rhein-treated and FTO-knockdown groups, adipocyte formation was impaired in both groups. Nonetheless, Rhein was shown to have m6A-independent targets, including the novel Receptor Expressing-Enhancing Protein 3 (REEP3). Huang group reports that REEP3, an endoplasmic reticulum (ER) regulator, promotes adipogenesis in an m6A-independent manner and represents a druggable candidate target for obesity therapeutics [117]. The aforementioned biochemical inhibitors reported by different research groups [91,111,112,114,115,116] provide insightful findings regarding FTO inhibition, to better understand its potential physiological function and clinical value. Yet, none has been critically evaluated in terms of their in-vivo efficacy in the context of FTO pathological implications like obesity or cancers.
A study in 2019 by Peng and colleagues [73] repurposed entacapone, an FDA-approved drug for treating Parkinson’s disease [118], as a potent chemical inhibitor of FTO. Their mechanistic findings showed that forkhead box protein O1 (FOXO1) mRNA is a downstream substrate of FTO, whereby entacapone inhibits FTO demethylase activity; resulting in m6A hypermethylated FOXO1 mRNA and reduced FOXO1 expression levels. A novel FTO-FOXO1 regulatory axis was identified using a high-fat diet-induced obese (DIO) mouse model, in which FTO inhibition via entacapone modulated hepatic gluconeogenesis as observed from lower fasting blood glucose levels, as well as modulating thermogenesis in adipose tissue as observed from increased UCP1 expression [73]. Nonetheless, the problematic high dosage used in animal studies and the differences between human and mice adipose tissue metabolism, necessitate the validation of entacapone use in humans and perhaps the development of entacapone analogs with better pharmacokinetic properties in the future.
A summary table of reported studies on biochemical inhibitors of FTO is illustrated in Table 2 below, highlighting current limitations as avenues for further future validations.

5. Concluding Remarks and Future Directions

Despite the discovery of RNA m6A in the 1970s [119,120] as the most prevalent modification found in eukaryotic mRNA; breakthroughs in better understanding its physiological implications began to be realized upon the identification of FTO as the first m6A demethylase of this highly dynamic and reversible process, as well as the development of m6A sequencing technologies. Dysfunctions in FTO m6A demethylase activity has been implicated in various human malignancies; highlighting m6A’s clinical role with insightful prognostic values for patients. In fact, given m6A regulatory role in various cancers, RNA methylation profiling can be utilized as a clinical tool in the future. The use of FTO inhibitors in conjunction with existing cancer therapies holds promising avenues for FTO-high expression tumors exhibiting resistance to cancer therapies as previously discussed [92,103]. FTO function in different cancers is context-dependent, whereby FTO may function as a tumor promoter or suppressor [106,107] by regulating various tissue-specific targets involved in unique pathways pertaining to different cancer types. Furthermore, taking into consideration that tumor angiogenesis is one of cancer microenvironment hallmarks [121] and that anti-angiogenic drug administration holds promise in improving treatment response [122]; further investigation on the involvement of m6A as an epitranscriptomic modification in cancer angiogenesis is needed. Current research on m6A involvement in cancer angiogenesis has mainly focused on an m6A writer METTL3 [123]; hence, future studies on other m6A regulators are required in the near future.
Several studies highlighted the underlying molecular mechanisms by which FTO is associated with obesity, given its role as an m6A demethylase, by denoting target gene transcripts that are affected by this post-transcriptional modification. Nevertheless, and as can be seen from Table 1, most adipogenesis studies have focused on FTO knockdown or overexpression in murine pre-adipocytes, which while advantageous does impose some limitations on adipogenesis studies. Since SNPs can impair function rather than complete loss of function (as performed in mice knockout studies); there is a need for more accurate models that reflect human FTO variants and their consequences in adipogenesis and obesity. Despite the benefits of mice 3T3-L1 pre-adipocytes in uncovering molecular mechanisms in adipogenesis; results obtained from human models are more reliable as they mimic human disorders, like obesity. In fact, primary human pre-adipocytes are utilized for the validation [124,125] or contradiction [124,126] of findings obtained using adipocyte animal models. Human primary pre-adipocytes are good tools for studying adipogenesis, and they can be derived from adipose tissue to produce in-vitro mature adipocytes under appropriate conditions. However, the drawbacks with these cells are their scarceness within fat tissue, and their limited renewal capacity [124]. Human-induced pluripotent stem cells (iPSCs) represent an attractive ideal cellular model that is promising in delineating mechanistic details of FTO SNPs on adipocyte differentiation, including production of thermogenic adipocytes at higher frequencies [127,128]. Several reports published directed differentiation protocols to produce thermogenic adipocytes using iPS cells without gene transfer methods [129,130,131,132]. Using iPS cells to produce thermogenic adipocytes and to investigate FTO’s role as an m6A demethylase throughout the adipogenesis process has not been investigated yet. This is a new and exciting area as it allows the temporal assessment of FTO and m6A during adipogenesis, including the earliest steps of adipogenesis such as adipocyte progenitors. This can bring about fundamental insights into adipogenesis regulation and the role of m6A in metabolic conditions like obesity.
Common pathways that are affected by FTO aberrant m6A regulatory functions in adipogenesis and tumorigenesis may likely exist, and ought to be further explored. Epidemiological studies reported strong associations of FTO SNPs with obesity, as well as various cancers. Some tumorigenic FTO SNPs were reported to be BMI-related [75,77,79] while others were independent of BMI [74,78,80,81]; indicating that obesity may confer increased risk in some cancers, through a potentially common FTO metabolic-tumorigenic regulatory axis. For instance, Liu and his team showed that FTO overexpression in breast cancer cells affects energy metabolism via the PI3K/AKT signaling pathway [86]. Autophagy might be a common regulatory mechanism in both disorders. Wang and colleagues report that FTO adipogenic regulatory role is exerted via ATG5 and ATG7-CEBPB signaling axis; delineating an underlying mechanism for the crosstalk between autophagy and adipogenesis [65]. A preliminary report by Zhao et al. pointed FTO’s oncogenic role in ovarian cancer, where FTO upregulation increased autophagy function of ovarian cancer cells, through increased autophagy-related 5 (ATG-5) protein levels [133]. Although this report is preliminary and lacks investigations on m6A levels, it points to the role of autophagy in tumorigenesis. Interestingly, the mTOR signaling pathway is a regulator of several cellular processes including autophagy, and is implicated in different disorders including cancer and obesity [134]. FTO-mediated tumorigenic role in endometrial cancer was uncovered and connected to the mTOR pathway [84,85,86], while mTOR’s connection to FTO-obesogenic functions awaits further exploration. In summary, molecular studies thus far are either focused on FTO molecular roles in adipogenesis or tumorigenesis, which will contribute to better understanding FTO molecular implications and awaits further investigations to decipher FTO’s potential dual function in obesity and cancer through common pathways.
Studies reviewed herein demonstrate that FTO is indeed an interesting druggable target in different human disorders, specifically cancers. Developing selective FTO inhibitors is still underway as studies investigating FTO inhibition have merely been tested in-silico context and some were tested in-vitro context. This necessitates the need for more in-vivo investigations as was performed in [73] which utilized a diet-induced obese (DIO) mouse model to investigate FTO inhibition in the context of obesity and adipogenesis. Performing in-vivo testing of FTO inhibitors is lagging; making it one of the current limitations in the field. Using relevant pre-clinical animal models can facilitate shifting these inhibitors into future clinical contexts. For example in the field of cancer therapeutics, patient-derived xenograft (PDX) models which were established by transplanting patient-specific tumor tissue into animals by surgery, were shown to maintain tumor microenvironment [135]. PDX models can be utilized as tools for personalized medicine strategies and cancer drug screening studies [135,136], such as testing FTO inhibitors in relation to its m6A-demethylase tumorigenic role.
As nearly one-third of mRNAs are m6A modified, the number of target genes regulated by FTO is relatively large. With such a large network of targeted transcripts reliant on the FTO enzymatic role, dysregulations in FTO m6A demethylase function must be further explored. Additional research into the biological implications of m6A modifications in various human pathologies, such as obesity and cancer, are required to fully understand their roles in such diseases and to develop precise therapies such as FTO small-molecule inhibitors in conjunction with traditional therapies, for better outcomes. Moving from GWAS signals to molecular and biological underpinnings necessitates the use of all technologies at our disposal, including third-generation sequencing technologies to highlight novel epitranscriptomic signatures in association with human disorders.

Author Contributions

Conceptualization; S.K.A. and A.A.S.; writing—original draft preparation, S.K.A.; writing—review and editing, S.K.A., H.A. and A.A.S.; project administration, A.A.S. and H.A.; funding acquisition, A.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Khalifa University of Science and Technology Competitive Internal Research Award CIRA-ADEK and by Abu Dhabi Award for Research Excellence (AARE) 2019.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gerken, T.; Girard, C.A.; Tung, Y.-C.L.; Webby, C.J.; Saudek, V.; Hewitson, K.S.; Yeo, G.S.H.; McDonough, M.A.; Cunliffe, S.; McNeill, L.A.; et al. The Obesity-Associated FTO Gene Encodes a 2-Oxoglutarate-Dependent Nucleic Acid Demethylase. Science 2007, 318, 1469–1472. [Google Scholar] [CrossRef] [Green Version]
  2. Jia, G.; Fu, Y.; Zhao, X.; Dai, Q.; Zheng, G.; Yang, Y.; Yi, C.; Lindahl, T.; Pan, T.; Yang, Y.-G.; et al. N6-Methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat. Chem. Biol. 2011, 7, 885–887. [Google Scholar] [CrossRef] [PubMed]
  3. Dominissini, D.; Moshitch-Moshkovitz, S.; Schwartz, S.; Salmon-Divon, M.; Ungar, L.; Osenberg, S.; Cesarkas, K.; Jacob-Hirsch, J.; Amariglio, N.; Kupiec, M.; et al. Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq. Nature 2012, 485, 201–206. [Google Scholar] [CrossRef] [PubMed]
  4. Mauer, J.; Luo, X.; Blanjoie, A.; Jiao, X.; Grozhik, A.V.; Patil, D.P.; Linder, B.; Pickering, B.F.; Vasseur, J.-J.; Chen, Q.; et al. Reversible methylation of m6Am in the 5′ cap controls mRNA stability. Nature 2017, 541, 371–375. [Google Scholar] [CrossRef] [Green Version]
  5. Anreiter, I.; Mir, Q.; Simpson, J.T.; Janga, S.C.; Soller, M. New Twists in Detecting mRNA Modification Dynamics. Trends Biotechnol. 2020, 39, 72–89. [Google Scholar] [CrossRef]
  6. Sajini, A.A.; Frye, M. RNA Methylation in the control of stem cell activity and epidermal differentiation. In Epigenetic Regulation of Skin Development and Regeneration (Stem Cell Biology and Regenerative Medicine); Botchkarev, V., Millar, S., Eds.; Humana Press: Cham, Switzerland, 2018; pp. 215–229. [Google Scholar] [CrossRef]
  7. Wang, X.; Lu, Z.; Gomez, A.; Hon, G.C.; Yue, Y.; Han, D.; Fu, Y.; Parisien, M.; Dai, Q.; Jia, G.; et al. N6-methyladenosine-dependent regulation of messenger RNA stability. Nature 2014, 505, 117–120. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, X.; Zhao, B.; Roundtree, I.A.; Lu, Z.; Han, D.; Ma, H.; Weng, X.; Chen, K.; Shi, H.; He, C. N6-methyladenosine Modulates Messenger RNA Translation Efficiency. Cell 2015, 161, 1388–1399. [Google Scholar] [CrossRef] [Green Version]
  9. Bartosovic, M.; Molares, H.C.; Gregorova, P.; Hrossova, D.; Kudla, G.; Vanacova, S. N6-methyladenosine demethylase FTO targets pre-mRNAs and regulates alternative splicing and 3′-end processing. Nucleic Acids Res. 2017, 45, 11356–11370. [Google Scholar] [CrossRef] [PubMed]
  10. Zhao, X.; Yang, Y.; Sun, B.-F.; Shi, Y.; Yang, X.; Xiao, W.; Hao, Y.-J.; Ping, X.-L.; Chen, Y.-S.; Wang, W.-J.; et al. FTO-dependent demethylation of N6-methyladenosine regulates mRNA splicing and is required for adipogenesis. Cell Res. 2014, 24, 1403–1419. [Google Scholar] [CrossRef] [PubMed]
  11. McElhinney, J.M.W.R.; Hasan, A.; Sajini, A.A. The epitranscriptome landscape of small noncoding RNAs in stem cells. Stem Cells 2020, 38, 1216–1228. [Google Scholar] [CrossRef] [PubMed]
  12. Ping, X.-L.; Sun, B.-F.; Wang, L.; Xiao, W.; Yang, X.; Wang, W.-J.; Adhikari, S.; Shi, Y.; Lv, Y.; Chen, Y.-S.; et al. Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Res. 2014, 24, 177–189. [Google Scholar] [CrossRef] [Green Version]
  13. Church, C.; Moir, L.; McMurray, F.; Girard, C.; Banks, G.; Teboul, L.; Wells, S.; Brüning, J.C.; Nolan, P.; Ashcroft, F.M.; et al. Overexpression of Fto leads to increased food intake and results in obesity. Nat. Genet. 2010, 42, 1086–1092. [Google Scholar] [CrossRef] [PubMed]
  14. Shen, F.; Huang, W.; Huang, J.-T.; Xiong, J.; Yang, Y.; Wu, K.; Jia, G.-F.; Chen, J.; Feng, Y.-Q.; Yuan, B.-F.; et al. DecreasedN6-Methyladenosine in Peripheral Blood RNA from Diabetic Patients Is Associated with FTO Expression rather than ALKBH5. J. Clin. Endocrinol. Metab. 2015, 100, E148–E154. [Google Scholar] [CrossRef] [Green Version]
  15. Mathiyalagan, P.; Adamiak, M.; Mayourian, J.; Sassi, Y.; Liang, Y.; Agarwal, N.; Jha, D.; Zhang, S.; Kohlbrenner, E.; Chepurko, E.; et al. FTO-Dependent N 6-Methyladenosine Regulates Cardiac Function during Remodeling and Repair. Circulation 2019, 139, 518–532. [Google Scholar] [CrossRef] [PubMed]
  16. Cui, Q.; Shi, H.; Ye, P.; Li, L.; Qu, Q.; Sun, G.; Sun, G.; Lu, Z.; Huang, Y.; Yang, C.-G.; et al. m6A RNA Methylation Regulates the Self-Renewal and Tumorigenesis of Glioblastoma Stem Cells. Cell Rep. 2017, 18, 2622–2634. [Google Scholar] [CrossRef] [PubMed]
  17. Han, M.; Liu, Z.; Xu, Y.; Liu, X.; Wang, D.; Li, F.; Wang, Y.; Bi, J. Abnormality of m6A mRNA Methylation Is Involved in Alzheimer’s Disease. Front. Neurosci. 2020, 14, 98. [Google Scholar] [CrossRef] [PubMed]
  18. Hess, M.E.; Hess, S.; Meyer, K.D.; Verhagen, L.A.W.; Koch, L.; Brönneke, H.S.; Dietrich, M.O.; Jordan, S.D.; Saletore, Y.; Elemento, O.; et al. The fat mass and obesity associated gene (Fto) regulates activity of the dopaminergic midbrain circuitry. Nat. Neurosci. 2013, 16, 1042–1048. [Google Scholar] [CrossRef] [PubMed]
  19. Frayling, T.M.; Timpson, N.J.; Weedon, M.N.; Zeggini, E.; Freathy, R.M.; Lindgren, C.M.; Perry, J.R.B.; Elliott, K.S.; Lango, H.; Rayner, N.W.; et al. A Common Variant in the FTO Gene Is Associated with Body Mass Index and Predisposes to Childhood and Adult Obesity. Science 2007, 316, 889–894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Scuteri, A.; Sanna, S.; Chen, W.-M.; Uda, M.; Albai, G.; Strait, J.; Najjar, S.; Nagaraja, R.; Orrú, M.; Usala, G.; et al. Genome-Wide Association Scan Shows Genetic Variants in the FTO Gene Are Associated with Obesity-Related Traits. PLoS Genet. 2007, 3, e115. [Google Scholar] [CrossRef] [PubMed]
  21. Dina, C.; Meyre, D.; Gallina, S.; Durand, E.; Körner, A.; Jacobson, P.; Carlsson, L.M.S.; Kiess, W.; Vatin, V.; Lecoeur, C.; et al. Variation in FTO contributes to childhood obesity and severe adult obesity. Nat. Genet. 2007, 39, 724–726. [Google Scholar] [CrossRef] [PubMed]
  22. Villalobos-Comparán, M.; Flores-Dorantes, M.T.; Molina, T.V.; Rodríguez-Cruz, M.; Ulloa, A.C.G.; Robles, L.; Huertas-Vázquez, A.; Saucedo-Villarreal, N.; López-Alarcón, M.; Sánchez-Muñoz, F.; et al. The FTO Gene Is Associated with Adulthood Obesity in the Mexican Population. Obesity 2008, 16, 2296–2301. [Google Scholar] [CrossRef] [PubMed]
  23. Dong, C.; Beecham, A.; Slifer, S.; Wang, L.; McClendon, M.S.; Blanton, S.H.; Rundek, T.; Sacco, R.L. Genome-wide linkage and peak-wide association study of obesity-related quantitative traits in Caribbean Hispanics. Qual. Life Res. 2011, 129, 209–219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Cho, Y.S.; Go, M.J.; Kim, Y.J.; Heo, J.Y.; Oh, J.H.; Ban, H.-J.; Yoon, D.; Lee, M.H.; Kim, D.-J.; Park, M.; et al. A large-scale genome-wide association study of Asian populations uncovers genetic factors influencing eight quantitative traits. Nat. Genet. 2009, 41, 527–534. [Google Scholar] [CrossRef]
  25. Wen, W.; Cho, Y.-S.; Zheng, W.; Dorajoo, R.; Kato, N.; Qi, L.; Chen, C.-H.; Delahanty, R.J.; Okada, Y. Meta-analysis identifies common variants associated with body mass index in east Asians. Nat. Genet. 2012, 44, 307–311. [Google Scholar] [CrossRef] [PubMed]
  26. Hotta, K.; Nakata, Y.; Matsuo, T.; Kamohara, S.; Kotani, K.; Komatsu, R.; Itoh, N.; Mineo, I.; Wada, J.; Masuzaki, H.; et al. Variations in the FTO gene are associated with severe obesity in the Japanese. J. Hum. Genet. 2008, 53, 546–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Rong, R.; Hanson, R.L.; Ortiz, D.; Wiedrich, C.; Kobes, S.; Knowler, W.C.; Bogardus, C.; Baier, L.J. Association Analysis of Variation in/Near FTO, CDKAL1, SLC30A8, HHEX, EXT2, IGF2BP2, LOC387761, and CDKN2B With Type 2 Diabetes and Related Quantitative Traits in Pima Indians. Diabetes 2009, 58, 478–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Rees, S.D.; Islam, M.; Hydrie, M.Z.I.; Chaudhary, B.; Bellary, S.; Hashmi, S.; O’Hare, J.P.; Kumar, S.; Sanghera, D.K.; Chaturvedi, N.; et al. An FTO variant is associated with Type 2 diabetes in South Asian populations after accounting for body mass index and waist circumference. Diabet. Med. 2011, 28, 673–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Vasan, S.K.; Fall, T.; Neville, M.J.; Antonisamy, B.; Fall, C.H.; Geethanjali, F.S.; Gu, H.F.; Raghupathy, P.; Samuel, P.; Thomas, N.; et al. Associations of Variants in FTO and Near MC4R with Obesity Traits in South Asian Indians. Obesity 2012, 20, 2268–2277. [Google Scholar] [CrossRef] [PubMed]
  30. Monda, K.L.; Chen, G.K.; Taylor, K.C.; Palmer, C.; Edwards, T.L.; Lange, L.A.; Ng, M.C.Y.; Adeyemo, A.A.; Allison, M.A.; Bielak, L.F.; et al. A meta-analysis identifies new loci associated with body mass index in individuals of African ancestry. Nat. Genet. 2013, 45, 690–696. [Google Scholar] [CrossRef] [Green Version]
  31. Khan, S.M.; Chehadeh, S.E.H.; Abdulrahman, M.; Osman, W.; Al Safar, H. Establishing a genetic link between FTO and VDR gene polymorphisms and obesity in the Emirati population. BMC Med. Genet. 2018, 19, 11. [Google Scholar] [CrossRef] [PubMed]
  32. Al-Serri, A.; Al-Bustan, S.A.; Kamkar, M.; Thomas, D.; Alsmadi, O.; Altemaimi, R.; Mojiminiyi, O.A.; Abdella, N.A. Association of FTO rs9939609 with Obesity in the Kuwaiti Population: A Public Health Concern? Med. Princ. Pract. 2018, 27, 145–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Ali, E.M.M.; Diab, T.; Elsaid, A.; El Daim, H.A.A.; Elshazli, R.M.; Settin, A. Fat mass and obesity-associated (FTO) and leptin receptor (LEPR) gene polymorphisms in Egyptian obese subjects. Arch. Physiol. Biochem. 2019, 3455, 28–36. [Google Scholar] [CrossRef] [PubMed]
  34. Cyrus, C.; Ismail, M.; Chathoth, S.; Vatte, C.; Hasen, M.; Al Ali, A. Analysis of the Impact of Common Polymorphisms of theFTOandMC4RGenes with the Risk of Severe Obesity in Saudi Arabian Population. Genet. Test. Mol. Biomark. 2018, 22, 170–177. [Google Scholar] [CrossRef] [PubMed]
  35. Osman, W.; Tay, G.; Alsafar, H. Multiple genetic variations confer risks for obesity and type 2 diabetes mellitus in Arab descendants from UAE. Int. J. Obes. 2018, 42, 1345–1353. [Google Scholar] [CrossRef] [PubMed]
  36. Chehadeh, S.E.H.; Osman, W.; Nazar, S.; Jerman, L.; Alghafri, A.; Sajwani, A.; Alawlaqi, M.; AlObeidli, M.; Jelinek, H.F.; AlAnouti, F.; et al. Implication of genetic variants in overweight and obesity susceptibility among the young Arab population of the United Arab Emirates. Gene 2020, 739, 144509. [Google Scholar] [CrossRef]
  37. Haupt, A.; Thamer, C.; Staiger, H.; Tschritter, O.; Kirchhoff, K.; Machicao, F.; Häring, H.-U.; Stefan, N.; Fritsche, A. Variation in the FTO Gene Influences Food Intake but not Energy Expenditure. Exp. Clin. Endocrinol. Diabetes 2009, 117, 194–197. [Google Scholar] [CrossRef]
  38. Cecil, J.E.; Ph, D.; Tavendale, R.; Ph, D.; Watt, P.; Ph, D.; Hetherington, M.M.; Ph, D.; Palmer, C.N.A.; Ph, D. An Obesity-Associated FTO Gene Variant and Increased Energy Intake in Children. N. Engl. J. Med. 2008, 359, 2558–2566. [Google Scholar] [CrossRef] [PubMed]
  39. Wardle, J.; Llewellyn, C.; Sanderson, S.; Plomin, R. The FTO gene and measured food intake in children. Int. J. Obes. 2009, 33, 42–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Wardle, J.; Carnell, S.; Haworth, C.; Farooqi, S.; O’Rahilly, S.; Plomin, R. Obesity Associated Genetic Variation in FTO is Associated with Diminished Satiety. J. Clin. Endocrinol. Metab. 2008, 93, 3640–3643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Speakman, J.R.; Rance, K.A.; Johnstone, A. Polymorphisms of the FTO Gene Are Associated with Variation in Energy Intake, but not Energy Expenditure. Obesity 2008, 16, 1961–1965. [Google Scholar] [CrossRef] [PubMed]
  42. Fischer, J.; Koch, L.; Emmerling, C.; Vierkotten, J.; Peters, T.; Brüning, J.C.; Rüther, U. Inactivation of the Fto gene protects from obesity. Nature 2009, 458, 894–898. [Google Scholar] [CrossRef] [PubMed]
  43. Church, C.; Lee, S.; Bagg, E.A.L.; McTaggart, J.S.; Deacon, R.; Gerken, T.; Lee, A.; Moir, L.; Mecinović, J.; Quwailid, M.M.; et al. A Mouse Model for the Metabolic Effects of the Human Fat Mass and Obesity Associated FTO Gene. PLoS Genet. 2009, 5, e1000599. [Google Scholar] [CrossRef] [Green Version]
  44. Gao, X.; Shin, Y.-H.; Li, M.; Wang, F.; Tong, Q.; Zhang, P. The Fat Mass and Obesity Associated Gene FTO Functions in the Brain to Regulate Postnatal Growth in Mice. PLoS ONE 2010, 5, e14005. [Google Scholar] [CrossRef] [Green Version]
  45. McMurray, F.; Church, C.D.; Larder, R.; Nicholson, G.; Wells, S.; Teboul, L.; Tung, Y.C.L.; Rimmington, D.; Bosch, F.; Jimenez, V.; et al. Adult Onset Global Loss of the Fto Gene Alters Body Composition and Metabolism in the Mouse. PLoS Genet. 2013, 9, e1003166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Merkestein, M.; Laber, S.; McMurray, F.; Andrew, D.; Sachse, G.; Sanderson, J.; Li, M.; Usher, S.; Sellayah, D.; Ashcroft, F.M.; et al. FTO influences adipogenesis by regulating mitotic clonal expansion. Nat. Commun. 2015, 6, 6792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Ronkainen, J.; Mondini, E.; Cinti, F.; Cinti, S.; Sebért, S.; Savolainen, M.J.; Salonurmi, T. Fto-Deficiency Affects the Gene and MicroRNA Expression Involved in Brown Adipogenesis and Browning of White Adipose Tissue in Mice. Int. J. Mol. Sci. 2016, 17, 1851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Yu, Q.; Liu, S.; Yu, L.; Xiao, Y.; Zhang, S.; Wang, X.; Xu, Y.; Yu, H.; Li, Y.; Yang, J.; et al. RNA demethylation increases the yield and biomass of rice and potato plants in field trials. Nat. Biotechnol. 2021, 39, 1581–1588. [Google Scholar] [CrossRef]
  49. Wang, X.; Zhu, L.; Chen, J.; Wang, Y. mRNA m6A methylation downregulates adipogenesis in porcine adipocytes. Biochem. Biophys. Res. Commun. 2015, 459, 201–207. [Google Scholar] [CrossRef]
  50. Boissel, S.; Reish, O.; Proulx, K.; Kawagoe-Takaki, H.; Sedgwick, B.; Yeo, G.S.H.; Meyre, D.; Golzio, C.; Molinari, F.; Kadhom, N.; et al. Loss-of-Function Mutation in the Dioxygenase-Encoding FTO Gene Causes Severe Growth Retardation and Multiple Malformations. Am. J. Hum. Genet. 2009, 85, 106–111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Daoud, H.; Zhang, D.; McMurray, F.; Yu, A.; Luco, S.M.; Vanstone, J.; Jarinova, O.; Carson, N.; Wickens, J.; Shishodia, S.; et al. Identification of a pathogenic FTO mutation by next-generation sequencing in a newborn with growth retardation and developmental delay. J. Med. Genet. 2015, 53, 200–207. [Google Scholar] [CrossRef] [PubMed]
  52. Karra, E.; Zelaya, F.O.; Rachel, L.; Karra, E.; Daly, O.G.O.; Choudhury, A.I.; Yousseif, A.; Millership, S.; Iwakura, H.; Akamizu, T.; et al. Food-cue responsivity Find the latest version: A link between FTO, ghrelin, and impaired brain food-cue responsivity. J. Clin. Investig. 2013, 123, 3539–3551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Tuck, M.T.; Wiehl, P.E.; Pan, T. Inhibition of 6-methyladenine formation decreases the translation efficiency of dihydrofolate reductase transcripts. Int. J. Biochem. Cell Biol. 1999, 31, 837–851. [Google Scholar] [CrossRef]
  54. Zhang, M.; Zhang, Y.; Ma, J.; Guo, F.; Cao, Q.; Zhang, Y.; Zhou, B.; Chai, J.; Zhao, W.; Zhao, R. The Demethylase Activity of FTO (Fat Mass and Obesity Associated Protein) Is Required for Preadipocyte Differentiation. PLoS ONE 2015, 10, e0133788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Han, Z.; Niu, T.; Chang, J.; Lei, X.; Zhao, M.; Wang, Q.; Cheng, W.; Wang, J.; Feng, Y.; Chai, J. Crystal structure of the FTO protein reveals basis for its substrate specificity. Nature 2010, 464, 1205–1209. [Google Scholar] [CrossRef]
  56. Stratigopoulos, G.; Carli, J.F.M.; O’Day, D.R.; Wang, L.; LeDuc, C.A.; Lanzano, P.; Chung, W.K.; Rosenbaum, M.; Egli, D.; Doherty, D.A.; et al. Hypomorphism for RPGRIP1L, a Ciliary Gene Vicinal to the FTO Locus, Causes Increased Adiposity in Mice. Cell Metab. 2015, 19, 767–779. [Google Scholar] [CrossRef] [Green Version]
  57. Smemo, S.; Tena, J.J.; Kim, K.-H.; Gamazon, E.R.; Sakabe, N.J.; Gómez-Marín, C.; Aneas, I.; Credidio, F.L.; Sobreira, D.R.; Wasserman, N.F.; et al. Obesity-associated variants within FTO form long-range functional connections with IRX3. Nature 2014, 507, 371–375. [Google Scholar] [CrossRef] [Green Version]
  58. Berulava, T.; Horsthemke, B. The obesity-associated SNPs in intron 1 of the FTO gene affect primary transcript levels. Eur. J. Hum. Genet. 2010, 18, 1054–1056. [Google Scholar] [CrossRef]
  59. Tews, D.; Fischer-Posovszky, P.; Fromme, T.; Klingenspor, M.; Fischer, J.; Rüther, U.; Marienfeld, R.; Barth, T.F.; Möller, P.; Debatin, K.-M.; et al. FTO Deficiency Induces UCP-1 Expression and Mitochondrial Uncoupling in Adipocytes. Endocrinology 2013, 154, 3141–3151. [Google Scholar] [CrossRef] [Green Version]
  60. Claussnitzer, M.; Dankel, S.N.; Kim, K.-H.; Quon, G.; Meuleman, W.; Haugen, C.; Glunk, V.; Sousa, I.S.; Beaudry, J.L.; Puviindran, V.; et al. FTO Obesity Variant Circuitry and Adipocyte Browning in Humans. N. Engl. J. Med. 2015, 373, 895–907. [Google Scholar] [CrossRef] [Green Version]
  61. Wu, W.; Feng, J.; Jiang, D.; Zhou, X.; Jiang, Q.; Cai, M.; Wang, X.; Shan, T.; Wang, Y. AMPK regulates lipid accumulation in skeletal muscle cells through FTO-dependent demethylation of N6-methyladenosine. Sci. Rep. 2017, 7, 41606. [Google Scholar] [CrossRef] [Green Version]
  62. Wu, R.; Liu, Y.; Yao, Y.; Zhao, Y.; Bi, Z.; Jiang, Q.; Liu, Q.; Cai, M.; Wang, F.; Wang, Y.; et al. FTO regulates adipogenesis by controlling cell cycle progression via m6A-YTHDF2 dependent mechanism. Biochim. Biophys. Acta BBA Mol. Cell Biol. Lipids 2018, 1863, 1323–1330. [Google Scholar] [CrossRef]
  63. Aguilar, V.; Fajas, L. Cycling through metabolism. EMBO Mol. Med. 2010, 2, 338–348. [Google Scholar] [CrossRef] [PubMed]
  64. Li, A.; Chen, Y.-S.; Ping, X.-L.; Yang, X.; Xiao, W.; Yang, Y.; Sun, H.-Y.; Zhu, Q.; Baidya, P.; Wang, X.; et al. Cytoplasmic m6A reader YTHDF3 promotes mRNA translation. Cell Res. 2017, 27, 444–447. [Google Scholar] [CrossRef] [Green Version]
  65. Song, T.; Yang, Y.; Wei, H.; Xie, X.; Lu, J.; Zeng, Q.; Peng, J.; Zhou, Y.; Jiang, S.; Peng, J. Zfp217 mediates m6A mRNA methylation to orchestrate transcriptional and post-transcriptional regulation to promote adipogenic differentiation. Nucleic Acids Res. 2019, 47, 6130–6144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Rooney, P.H.; Boonsong, A.; McFadyen, M.C.; McLeod, H.L.; Cassidy, J.; Curran, S.; I Murray, G. The candidate oncogeneZNF217 is frequently amplified in colon cancer. J. Pathol. 2004, 204, 282–288. [Google Scholar] [CrossRef]
  67. Quinlan, K.G.; Verger, A.; Yaswen, P.; Crossley, M. Amplification of zinc finger gene 217 (ZNF217) and cancer: When good fingers go bad. Biochim. Biophys. Acta 2007, 1775, 333–340. [Google Scholar] [CrossRef] [Green Version]
  68. Rahman, M.T.; Nakayama, K.; Rahman, M.; Nakayama, N.; Ishikawa, M.; Katagiri, A.; Iida, K.; Nakayama, S.; Otsuki, Y.; Shih, I.-M.; et al. Prognostic and therapeutic impact of the chromosome 20q13.2 ZNF217 locus amplification in ovarian clear cell carcinoma. Cancer 2012, 118, 2846–2857. [Google Scholar] [CrossRef] [Green Version]
  69. Xiang, H.; Zhong, Z.-X.; Peng, Y.-D.; Jiang, S.-W. The Emerging Role of Zfp217 in Adipogenesis. Int. J. Mol. Sci. 2017, 18, 1367. [Google Scholar] [CrossRef]
  70. Wang, L.; Song, C.; Wang, N.; Li, S.; Liu, Q.; Sun, Z.; Wang, K.; Yu, S.C.; Yang, Q. NADP Modulates RNA m6A Methylation and Adipogenesis via Enhancing FTO Activity. Nat. Chem. Biol. 2020, 16, 1394–1402. [Google Scholar] [CrossRef]
  71. Wang, X.; Wu, R.; Liu, Y.; Zhao, Y.; Bi, Z.; Yao, Y.; Liu, Q.; Shi, H.; Wang, F.; Wang, Y. m6A mRNA methylation controls autophagy and adipogenesis by targeting Atg5 and Atg7. Autophagy 2020, 16, 1221–1235. [Google Scholar] [CrossRef]
  72. Wu, R.; Chen, Y.; Liu, Y.; Zhuang, L.; Chen, W.; Zeng, B.; Liao, X.; Guo, G.; Wang, Y.; Wang, X. m6A methylation promotes white-to-beige fat transition by facilitating Hif1a translation. EMBO Rep. 2021, 22, e52348. [Google Scholar] [CrossRef] [PubMed]
  73. Peng, S.; Xiao, W.; Ju, D.; Sun, B.; Hou, N.; Liu, Q.; Wang, Y.; Zhao, H.; Gao, C.; Zhang, S.; et al. Identification of entacapone as a chemical inhibitor of FTO mediating metabolic regulation through FOXO1. Sci. Transl. Med. 2019, 11, eaau7116. [Google Scholar] [CrossRef]
  74. Delahanty, R.J.; Beeghly-Fadiel, A.; Xiang, Y.-B.; Long, J.; Cai, Q.; Wen, W.; Xu, W.; Cai, H.; He, J.; Gao, Y.-T.; et al. Association of Obesity-related Genetic Variants with Endometrial Cancer Risk: A Report from the Shanghai Endometrial Cancer Genetics Study. Am. J. Epidemiol. 2011, 174, 1115–1126. [Google Scholar] [CrossRef] [PubMed]
  75. Lurie, G.; Gaudet, M.M.; Spurdle, A.; Carney, M.E.; Wilkens, L.R.; Yang, H.P.; Weiss, N.S.; Webb, P.; Thompson, P.J.; Terada, K.; et al. The Obesity-Associated Polymorphisms FTO rs9939609 and MC4R rs17782313 and Endometrial Cancer Risk in Non-Hispanic White Women. PLoS ONE 2011, 6, e16756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Huang, X.; Zhao, J.; Yang, M.; Li, M.; Zheng, J. Association between FTO gene polymorphism (rs9939609 T/A) and cancer risk: A meta-analysis. Eur. J. Cancer Care 2017, 26, e12464. [Google Scholar] [CrossRef]
  77. Kaklamani, V.; Yi, N.; Sadim, M.; Siziopikou, K.; Zhang, K.; Xu, Y.; Tofilon, S.; Agarwal, S.; Pasche, B.; Mantzoros, C. The role of the fat mass and obesity associated gene (FTO) in breast cancer risk. BMC Med. Genet. 2011, 12, 52. [Google Scholar] [CrossRef] [Green Version]
  78. Garcia-Closas, M.; Couch, F.J.; Lindstrom, S.; Michailidou, K.; Schmidt, M.K.; Brook, M.N.; Orr, N.; Rhie, S.K.; Riboli, E.; Feigelson, H.S.; et al. Genome-wide association studies identify four ER negative–specific breast cancer risk loci. Nat. Genet. 2013, 45, 392–398. [Google Scholar] [CrossRef] [Green Version]
  79. Tang, H.; Dong, X.; Hassan, M.; Abbruzzese, J.L.; Li, D. Body Mass Index and Obesity- and Diabetes-Associated Genotypes and Risk for Pancreatic Cancer. Cancer Epidemiol. Biomark. Prev. 2011, 20, 779–792. [Google Scholar] [CrossRef] [Green Version]
  80. Lin, Y.; Ueda, J.; Yagyu, K.; Ishii, H.; Ueno, M.; Egawa, N.; Nakao, H.; Mori, M.; Matsuo, K.; Kikuchi, S. Association between variations in the fat mass and obesity-associated gene and pancreatic cancer risk: A case–control study in Japan. BMC Cancer 2013, 13, 1. [Google Scholar] [CrossRef] [Green Version]
  81. Iles, M.M.; Law, M.H.; Stacey, S.N.; Han, J.; Fang, S.; Pfeiffer, R.; Harland, M.; MacGregor, S.; Taylor, J.C.; Aben, K.K.; et al. A variant in FTO shows association with melanoma risk not due to BMI. Nat. Genet. 2013, 45, 428–432. [Google Scholar] [CrossRef] [Green Version]
  82. Zhang, Z.; Zhou, D.; Lai, Y.; Liu, Y.; Tao, X.; Wang, Q.; Zhao, G.; Gu, H.; Liao, H.; Zhu, Y.; et al. Estrogen induces endometrial cancer cell proliferation and invasion by regulating the fat mass and obesity-associated gene via PI3K/AKT and MAPK signaling pathways. Cancer Lett. 2012, 319, 89–97. [Google Scholar] [CrossRef] [PubMed]
  83. Tan, A.; Dang, Y.; Chen, G.; Mo, Z. Overexpression of the fat mass and obesity associated gene (FTO) in breast cancer and its clinical implications. Int. J. Clin. Exp. Pathol. 2015, 8, 13405–13410. [Google Scholar] [PubMed]
  84. Zhu, Y.; Shen, J.; Gao, L.; Feng, Y. Estrogen promotes fat mass and obesity-associated protein nuclear localization and enhances endometrial cancer cell proliferation via the mTOR signaling pathway. Oncol. Rep. 2016, 35, 2391–2397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Bradford, L.S.; Rauh-Hain, A.; Clark, R.M.; Groeneweg, J.W.; Zhang, L.; Borger, D.; Zukerberg, L.R.; Growdon, W.B.; Foster, R.; Rueda, B.R. Assessing the efficacy of targeting the phosphatidylinositol 3-kinase/AKT/mTOR signaling pathway in endometrial cancer. Gynecol. Oncol. 2014, 133, 346–352. [Google Scholar] [CrossRef]
  86. Liu, Y.; Wang, R.; Zhang, L.; Li, J.; Lou, K.; Shi, B. The lipid metabolism gene FTO influences breast cancer cell energy metabolism via the PI3K/AKT signaling pathway. Oncol. Lett. 2017, 13, 4685–4690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Vivanco, I.; Sawyers, C.L. The phosphatidylinositol 3-Kinase–AKT pathway in human cancer. Nat. Cancer 2002, 2, 489–501. [Google Scholar] [CrossRef]
  88. Luo, J.; Manning, B.D.; Cantley, L.C. Targeting the PI3K-Akt pathway in human cancer: Rationale and promise. Cancer Cell 2003, 4, 257–262. [Google Scholar] [CrossRef] [Green Version]
  89. Xu, D.; Shao, W.; Jiang, Y.; Wang, X.; Liu, Y.; Liu, X. FTO expression is associated with the occurrence of gastric cancer and prognosis. Oncol. Rep. 2017, 38, 2285–2292. [Google Scholar] [CrossRef] [Green Version]
  90. Li, Z.; Weng, H.; Su, R.; Weng, X.; Zuo, Z.; Li, C.; Huang, H.; Nachtergaele, S.; Dong, L.; Hu, C.; et al. FTO Plays an Oncogenic Role in Acute Myeloid Leukemia as a N 6 -Methyladenosine RNA Demethylase. Cancer Cell 2017, 31, 127–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Huang, Y.; Yan, J.; Li, Q.; Li, J.; Gong, S.; Zhou, H.; Gan, J.; Jiang, H.; Jia, G.-F.; Luo, C.; et al. Meclofenamic acid selectively inhibits FTO demethylation of m6A over ALKBH5. Nucleic Acids Res. 2015, 43, 373–384. [Google Scholar] [CrossRef]
  92. Su, R.; Dong, L.; Li, C.; Nachtergaele, S.; Wunderlich, M.; Qing, Y.; Deng, X.; Wang, Y.; Weng, X.; Hu, C.; et al. R-2HG Exhibits Anti-tumor Activity by Targeting FTO/m6A/MYC/CEBPA Signaling. Cell 2018, 172, 90–105. [Google Scholar] [CrossRef] [Green Version]
  93. Figueroa, M.E.; Abdel-Wahab, O.; Lu, C.; Ward, P.S.; Patel, J.; Shih, A.; Li, Y.; Bhagwat, N.; VasanthaKumar, A.; Fernandez, H.F.; et al. Leukemic IDH1 and IDH2 Mutations Result in a Hypermethylation Phenotype, Disrupt TET2 Function, and Impair Hematopoietic Differentiation. Cancer Cell 2010, 18, 553–567. [Google Scholar] [CrossRef] [Green Version]
  94. Xu, W.; Yang, H.; Liu, Y.; Yang, Y.; Wang, P.; Kim, S.-H.; Ito, S.; Yang, C.; Wang, P.; Xiao, M.-T.; et al. Oncometabolite 2-Hydroxyglutarate Is a Competitive Inhibitor of α-Ketoglutarate-Dependent Dioxygenases. Cancer Cell 2011, 19, 17–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zhou, S.; Bai, Z.-L.; Xia, D.; Zhao, Z.-J.; Zhao, R.; Wang, Y.-Y.; Zhe, H. FTO regulates the chemo-radiotherapy resistance of cervical squamous cell carcinoma (CSCC) by targeting β-catenin through mRNA demethylation. Mol. Carcinog. 2018, 57, 590–597. [Google Scholar] [CrossRef] [PubMed]
  96. Zou, D.; Dong, L.; Li, C.; Yin, Z.; Rao, S.; Zhou, Q. The m6A eraser FTO facilitates proliferation and migration of human cervical cancer cells. Cancer Cell Int. 2019, 19, 321. [Google Scholar] [CrossRef] [Green Version]
  97. Liu, J.; Ren, D.; Du, Z.; Wang, H.; Zhang, H.; Jin, Y. m6A demethylase FTO facilitates tumor progression in lung squamous cell carcinoma by regulating MZF1 expression. Biochem. Biophys. Res. Commun. 2018, 502, 456–464. [Google Scholar] [CrossRef]
  98. Li, J.; Han, Y.; Zhang, H.; Qian, Z.; Jia, W.; Gao, Y.; Zheng, H.; Li, B. The m6A demethylase FTO promotes the growth of lung cancer cells by regulating the m6A level of USP7 mRNA. Biochem. Biophys. Res. Commun. 2019, 512, 479–485. [Google Scholar] [CrossRef]
  99. Tsuruta, N.; Tsuchihashi, K.; Ohmura, H.; Yamaguchi, K.; Ito, M.; Ariyama, H.; Kusaba, H.; Akashi, K.; Baba, E. RNA N6-methyladenosine demethylase FTO regulates PD-L1 expression in colon cancer cells. Biochem. Biophys. Res. Commun. 2020, 530, 235–239. [Google Scholar] [CrossRef]
  100. Zou, W.; Wolchok, J.D.; Chen, L. PD-L1 (B7-H1) and PD-1 pathway blockade for cancer therapy: Mechanisms, response biomarkers, and combinations. Sci. Transl. Med. 2016, 8, 328rv4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Niu, Y.; Lin, Z.; Wan, A.; Chen, H.; Liang, H.; Sun, L.; Wang, Y.; Li, X.; Xiong, X.-F.; Wei, B.; et al. RNA N6-methyladenosine demethylase FTO promotes breast tumor progression through inhibiting BNIP3. Mol. Cancer 2019, 18, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Cantor, S.; Xie, J.; Brandau, S.; Irminger-Finger, I.; Toro, J.R.; Bale, S.; Hamouda, R.S.; Cryns, V.L.; Jumppanen, M.; Isola, J.; et al. Bcl-2 Family Proteins. Encycl. Cancer 2011, 361. [Google Scholar] [CrossRef]
  103. Yang, S.; Wei, J.; Cui, Y.-H.; Park, G.; Shah, P.; Deng, Y.; Aplin, A.E.; Lu, Z.; Hwang, S.; He, C.; et al. m6A mRNA demethylase FTO regulates melanoma tumorigenicity and response to anti-PD-1 blockade. Nat. Commun. 2019, 10, 2782. [Google Scholar] [CrossRef]
  104. Huang, Y.; Su, R.; Sheng, Y.; Dong, L.; Dong, Z.; Xu, H.; Ni, T.; Zhang, Z.; Zhang, T.; Li, C.; et al. Small-Molecule Targeting of Oncogenic FTO Demethylase in Acute Myeloid Leukemia. Cancer Cell 2019, 35, 677–691.e10. [Google Scholar] [CrossRef] [Green Version]
  105. Zhang, L.; Wan, Y.; Zhang, Z.; Jiang, Y.; Lang, J.; Cheng, W.; Zhu, L. FTO demethylates m6A modifications in HOXB13 mRNA and promotes endometrial cancer metastasis by activating the WNT signalling pathway. RNA Biol. 2021, 18, 1265–1278. [Google Scholar] [CrossRef] [PubMed]
  106. Zeng, J.; Zhang, H.; Tan, Y.; Wang, Z.; Li, Y.; Yang, X. m6A demethylase FTO suppresses pancreatic cancer tumorigenesis by demethylating PJA2 and inhibiting Wnt signaling. Mol. Ther-Nucleic Acids 2021, 25, 277–292. [Google Scholar] [CrossRef] [PubMed]
  107. Tang, X.; Liu, S.; Chen, D.; Zhao, Z.; Zhou, J. The role of the fat mass and obesity-associated protein in the proliferation of pancreatic cancer cells. Oncol. Lett. 2019, 17, 2473–2478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Bretones, G.; Delgado, M.D.; León, J. Myc and cell cycle control. Biochim. Biophys. Acta BBA Gene Regul. Mech. 2015, 1849, 506–516. [Google Scholar] [CrossRef] [PubMed]
  109. Huang, H.; Wang, Y.; Kandpal, M.; Zhao, G.; Cardenas, H.; Ji, Y.; Chaparala, A.; Tanner, E.J.; Chen, J.; Davuluri, R.V.; et al. FTO-Dependent N6-Methyladenosine Modifications Inhibit Ovarian Cancer Stem Cell Self-Renewal by Blocking cAMP Signaling. Cancer Res. 2020, 80, 3200–3214. [Google Scholar] [CrossRef]
  110. Huang, H.; Weng, H.; Sun, W.; Qin, X.; Shi, H.; Wu, H.; Zhao, B.S.; Mesquita, A.; Liu, C.; Yuan, C.L.; et al. Recognition of RNA N6-methyladenosine by IGF2BP proteins enhances mRNA stability and translation. Nat. Cell Biol. 2018, 20, 285–295. [Google Scholar] [CrossRef] [PubMed]
  111. Chen, B.; Ye, F.; Yu, L.; Jia, G.; Huang, X.; Zhang, X.; Peng, S.; Chen, K.; Wang, M.; Gong, S.; et al. Development of Cell-Active N6-Methyladenosine RNA Demethylase FTO Inhibitor. J. Am. Chem. Soc. 2012, 134, 17963–17971. [Google Scholar] [CrossRef]
  112. Aik, W.S.; Demetriades, M.; Hamdan, M.K.K.; Bagg, E.; Yeoh, K.K.; Lejeune, C.; Zhang, Z.; McDonough, M.A.; Schofield, C.J. Structural Basis for Inhibition of the Fat Mass and Obesity Associated Protein (FTO). J. Med. Chem. 2013, 56, 3680–3688. [Google Scholar] [CrossRef]
  113. Bartzatt, R. Anti-Inflammatory Drugs and Prediction of New Structures by Comparative Analysis. Anti-Inflamm. Anti-Allergy Agents Med. Chem. 2012, 11, 151–160. [Google Scholar] [CrossRef]
  114. Zheng, G.; Cox, T.; Tribbey, L.; Wang, G.Z.; Iacoban, P.; Booher, M.E.; Gabriel, G.J.; Zhou, L.; Bae, N.; Rowles, J.; et al. Synthesis of a FTO Inhibitor with Anticonvulsant Activity. ACS Chem. Neurosci. 2014, 5, 658–665. [Google Scholar] [CrossRef] [PubMed]
  115. Toh, J.D.W.; Sun, L.; Lau, L.Z.M.; Tan, J.; Low, J.J.A.; Tang, C.W.Q.; Cheong, E.J.Y.; Tan, M.J.H.; Chen, Y.; Hong, W.; et al. A strategy based on nucleotide specificity leads to a subfamily-selective and cell-active inhibitor of N6-methyladenosine demethylase FTO. Chem. Sci. 2015, 6, 112–122. [Google Scholar] [CrossRef]
  116. He, W.; Zhou, B.; Liu, W.; Zhang, M.; Shen, Z.; Han, Z.; Jiang, Q.; Yang, Q.; Song, C.; Wang, R.; et al. Identification of A Novel Small-Molecule Binding Site of the Fat Mass and Obesity Associated Protein (FTO). J. Med. Chem. 2015, 58, 7341–7348. [Google Scholar] [CrossRef]
  117. Huang, L.; Zhang, J.; Zhu, X.; Mi, X.; Li, Q.; Gao, J.; Zhou, J.; Zhou, J.; Liu, X.-M. The Phytochemical Rhein Mediates M6A-Independent Suppression of Adipocyte Differentiation. Front. Nutr. 2021, 8, 831. [Google Scholar] [CrossRef]
  118. Nissinen, E.; Linden, I.-B.; Schultz, E.; Pohto, P. Biochemical and pharmacological properties of a peripherally acting catechol-O-methyltransferase inhibitor entacapone. Naunyn-Schmiedeberg’s Arch. Pharmacol. 1992, 346, 262–266. [Google Scholar] [CrossRef]
  119. Desrosiers, R.; Friderici, K.; Rottman, F. Identification of Methylated Nucleosides in Messenger RNA from Novikoff Hepatoma Cells. Proc. Natl. Acad. Sci. USA 1974, 71, 3971–3975. [Google Scholar] [CrossRef] [Green Version]
  120. Perry, R.P.; Kelley, D.E. Existence of methylated messenger RNA in mouse L cells. Cell 1974, 1, 37–42. [Google Scholar] [CrossRef]
  121. Solimando, A.G.; de Summa, S.; Vacca, A.; Ribatti, D. Cancer-Associated Angiogenesis: The Endothelial Cell as a Checkpoint for Immunological Patrolling. Cancers 2020, 12, 3380. [Google Scholar] [CrossRef]
  122. Ribatti, D.; Solimando, A.; Pezzella, F. The Anti-VEGF(R) Drug Discovery Legacy: Improving Attrition Rates by Breaking the Vicious Cycle of Angiogenesis in Cancer. Cancers 2021, 13, 3433. [Google Scholar] [CrossRef] [PubMed]
  123. Li, H.; Wu, H.; Wang, Q.; Ning, S.; Xu, S.; Pang, D. Dual effects of N6-methyladenosine on cancer progression and immunotherapy. Mol. Ther. Nucleic Acids 2021, 24, 25–39. [Google Scholar] [CrossRef] [PubMed]
  124. Ruiz-Ojeda, F.J.; Rupérez, A.I.; Gomez-Llorente, C.; Gil, A.; Aguilera, C.M. Cell Models and Their Application for Studying Adipogenic Differentiation in Relation to Obesity: A Review. Int. J. Mol. Sci. 2016, 17, 1040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Todoric, J.; Strobl, B.; Jais, A.; Boucheron, N.; Bayer, M.; Amann, S.; Lindroos, J.; Teperino, R.; Prager, G.; Bilban, M.; et al. Cross-Talk between Interferon-γ and Hedgehog Signaling Regulates Adipogenesis. Diabetes 2011, 60, 1668–1676. [Google Scholar] [CrossRef] [Green Version]
  126. Gagnon, A.; Foster, C.; Landry, A.; Sorisky, A. The role of interleukin 1β in the anti-adipogenic action of macrophages on human preadipocytes. J. Endocrinol. 2013, 217, 197–206. [Google Scholar] [CrossRef] [Green Version]
  127. Yao, X.; Dani, V.; Dani, C. Human Pluripotent Stem Cells: A Relevant Model to Identify Pathways Governing Thermogenic Adipocyte Generation. Front. Endocrinol. 2020, 10, 932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Bahmad, H.F.; Daouk, R.; Azar, J.; Sapudom, J.; Teo, J.C.M.; Abou-Kheir, W.; Al-Sayegh, M. Modeling Adipogenesis: Current and Future Perspective. Cells 2020, 9, 2326. [Google Scholar] [CrossRef]
  129. Nishio, M.; Yoneshiro, T.; Nakahara, M.; Suzuki, S.; Saeki, K.; Hasegawa, M.; Kawai, Y.; Akutsu, H.; Umezawa, A.; Yasuda, K.; et al. Production of Functional Classical Brown Adipocytes from Human Pluripotent Stem Cells using Specific Hemopoietin Cocktail without Gene Transfer. Cell Metab. 2012, 16, 394–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Guénantin, A.-C.; Briand, N.; Capel, E.; Dumont, F.; Morichon, R.; Provost, C.; Stillitano, F.; Jeziorowska, D.; Siffroi, J.-P.; Hajjar, R.J.; et al. Functional Human Beige Adipocytes from Induced Pluripotent Stem Cells. Diabetes 2017, 66, 1470–1478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Su, S.; Guntur, A.R.; Nguyen, D.C.; Fakory, S.S.; Doucette, C.C.; Leech, C.; Lotana, H.; Kelley, M.; Kohli, J.; Martino, J.; et al. A Renewable Source of Human Beige Adipocytes for Development of Therapies to Treat Metabolic Syndrome. Cell Rep. 2018, 25, 3215–3228.e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Carobbio, S.; Guenantin, A.-C.; Bahri, M.; Rodriguez-Fdez, S.; Honig, F.; Kamzolas, I.; Samuelson, I.; Long, K.; Awad, S.; Lukovic, D.; et al. Unraveling the Developmental Roadmap toward Human Brown Adipose Tissue. Stem Cell Rep. 2021, 16, 641–655. [Google Scholar] [CrossRef] [PubMed]
  133. Zhao, L.; Kong, X.; Zhong, W.; Wang, Y.; Li, P. FTO accelerates ovarian cancer cell growth by promoting proliferation, inhibiting apoptosis, and activating autophagy. Pathol. Res. Pract. 2020, 216, 153042. [Google Scholar] [CrossRef] [PubMed]
  134. Laplante, M.; Sabatini, D.M. mTOR Signaling in Growth Control and Disease. Cell 2012, 149, 274–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Li, Z.; Zheng, W.; Wang, H.; Cheng, Y.; Fang, Y.; Wu, F.; Sun, G.; Sun, G.; Lv, C.; Hui, B. Application of Animal Models in Cancer Research: Recent Progress and Future Prospects. Cancer Manag. Res. 2021, 13, 2455–2475. [Google Scholar] [CrossRef]
  136. Hidalgo, M.; Amant, F.; Biankin, A.V.; Budinská, E.; Byrne, A.T.; Caldas, C.; Clarke, R.B.; de Jong, S.; Jonkers, J.; Mælandsmo, G.M.; et al. Patient-Derived Xenograft Models: An Emerging Platform for Translational Cancer Research. Cancer Discov. 2014, 4, 998–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. FTO molecular mechanisms in association with adipogenic pathways. (A) Inverse correlation between FTO overexpression and lipid accumulation as reported in porcine adipocytes. (B) Overexpression of FTO in human subjects homozygous for FTO SNP rs9939609 risk allele, resulting in increased levels of the hunger hormone ghrelin. (C) m6A and alternative splicing regulation: perturbed binding of SRSF2 to RUNX1T1 results in the skipping exon 6 and generating the pro-adipogenic RUNX1T1 S variant. (D) Mechanistically, RUNX1T1 S-isoform promotes adipogenic differentiation via increasing D-type Cyclin genes (Cyclin D1 & D3) during the MCE phase. (E) FTO regulatory role in skeletal muscle cells’ lipid accumulation capacity: AMPK activation downregulates FTO, resulting in reduced lipid accumulation. (F) FTO regulatory role in adipocyte cell cycle progression: FTO inhibition results in m6A hyper-methylation of two mitotic regulators’ transcripts CCNA2 and CDK2, which are recognized and degraded by YTHDF2 m6A reader; impairing cell cycle and suppressing adipogenesis. (G) Zfp217 interaction with FTO promotes adipogenesis via transcriptional and post-transcriptional regulatory mechanisms. (H) Association between metabolism and RNA m6A demethylation: NADP binding to FTO enhances its activity and promotes adipogenesis. (I) Connecting m6A role in adipogenesis with autophagy: FTO reduces m6A levels on two autophagy-related genes ATG5 and ATG7, stabilizing them from decay by YTHDF2; thus promoting autophagy and adipogenesis. (J) FTO promotes thermogenesis and white-to-beige fat transition: FTO knockdown produces m6A hyper-methylated HIF1A and increases its translation through a YTHDC2-mediated process; HIF1A, in turn, activates thermogenic genes PGC1A, PRDM16, and PPARG, which promote white adipocyte “browning”, as an anti-obesity approach. Abbreviations: RUNX1T1: Runt-related transcription factor 1; SRSF2: splicing regulatory protein; MCE: mitotic clonal expansion phase; AMPK: AMP-activated protein kinase; CCNA2: cyclin A2; CDK2: cyclin-dependent kinase 2; Zfp217: Zinc finger protein 217; NADP: nicotinamide adenine dinucleotide phosphate; ATG5: autophagy-related gene 5; ATG7: autophagy-related gene 7; HIF1A: hypoxia inducible factor 1 subunit alpha.
Figure 1. FTO molecular mechanisms in association with adipogenic pathways. (A) Inverse correlation between FTO overexpression and lipid accumulation as reported in porcine adipocytes. (B) Overexpression of FTO in human subjects homozygous for FTO SNP rs9939609 risk allele, resulting in increased levels of the hunger hormone ghrelin. (C) m6A and alternative splicing regulation: perturbed binding of SRSF2 to RUNX1T1 results in the skipping exon 6 and generating the pro-adipogenic RUNX1T1 S variant. (D) Mechanistically, RUNX1T1 S-isoform promotes adipogenic differentiation via increasing D-type Cyclin genes (Cyclin D1 & D3) during the MCE phase. (E) FTO regulatory role in skeletal muscle cells’ lipid accumulation capacity: AMPK activation downregulates FTO, resulting in reduced lipid accumulation. (F) FTO regulatory role in adipocyte cell cycle progression: FTO inhibition results in m6A hyper-methylation of two mitotic regulators’ transcripts CCNA2 and CDK2, which are recognized and degraded by YTHDF2 m6A reader; impairing cell cycle and suppressing adipogenesis. (G) Zfp217 interaction with FTO promotes adipogenesis via transcriptional and post-transcriptional regulatory mechanisms. (H) Association between metabolism and RNA m6A demethylation: NADP binding to FTO enhances its activity and promotes adipogenesis. (I) Connecting m6A role in adipogenesis with autophagy: FTO reduces m6A levels on two autophagy-related genes ATG5 and ATG7, stabilizing them from decay by YTHDF2; thus promoting autophagy and adipogenesis. (J) FTO promotes thermogenesis and white-to-beige fat transition: FTO knockdown produces m6A hyper-methylated HIF1A and increases its translation through a YTHDC2-mediated process; HIF1A, in turn, activates thermogenic genes PGC1A, PRDM16, and PPARG, which promote white adipocyte “browning”, as an anti-obesity approach. Abbreviations: RUNX1T1: Runt-related transcription factor 1; SRSF2: splicing regulatory protein; MCE: mitotic clonal expansion phase; AMPK: AMP-activated protein kinase; CCNA2: cyclin A2; CDK2: cyclin-dependent kinase 2; Zfp217: Zinc finger protein 217; NADP: nicotinamide adenine dinucleotide phosphate; ATG5: autophagy-related gene 5; ATG7: autophagy-related gene 7; HIF1A: hypoxia inducible factor 1 subunit alpha.
Ijms 23 03800 g001
Figure 2. FTO upregulation in various human cancers and its molecular implications as an m6A demethylase, indicated by the regulation of different targets and affected signaling pathways. (A) FTO upregulation in AML modulates ASB2/RARA levels, hence, promoting AML cellular viability and proliferation. FTO inhibition via FB23-2 inhibits AML progression, by upregulating ASB2/RARA and downregulating MYC/CEBPA. (B) Given FTO’s tumorigenic role in glioblastoma, its inhibition via MA2 suppresses GSCs self-renewal and tumorigenesis. Additionally, inhibiting FTO using R-2HG exerts anti-tumor functions in lethal glioma and leukemia by modulating MYC/CEBPA levels. (C) FTO upregulation in CSCC induces chemo-radiotherapy resistance by increasing β-catenin and activating the ERCC1 pathway. E2F1/MYC are also upregulated targets in CSCC that promote oncogenic functions. (D) FTO overexpression in lung cancer upregulates MZF1 and USP7 levels via modulating their m6A levels. Thus, promoting lung cancer cells’ proliferation, invasion, and colony formation, while inhibiting their apoptosis. (E) In breast cancer cells, FTO upregulation demethylates BNIP3 as a downstream target and downregulates it in a YTHDF2-independent manner, and promoting oncogenic roles like metastasis. (F) In melanoma, upregulated FTO and reduced m6A levels, stabilize melanoma-promoting genes namely, PDCD1, CXCR4, and SOX10 in a YTHDF2-mediated process. Consequently, increasing melanoma migration, proliferation, and immunotherapy resistance. (G) In endometrial carcinoma, higher HOXB13 expression activates Wnt signaling and promotes metastasis. (H) In pancreatic cancer, FTO overexpression leads to upregulating c-MYC oncogene and enhancing pancreatic cancer proliferation. FTO’s additional role as a tumor suppressor in pancreatic cancer is discussed in the text. Abbreviations: ASB2: Ankyrin repeat and SOCS box protein 2; RARA: retinoic acid receptor α; GSCs: glioblastoma stem cells; MA: Meclofenamic acid; R-2HG: R-2-hydroxyglutarate; ERCC1: excision repair cross-complementation group 1; MZF1: Myeloid Zinc Finger Protein 1; USP7: ubiquitin-specific protease-7; BNIP3: BCL2 Interacting Protein 3; HOXB13: homeobox transcription factor.
Figure 2. FTO upregulation in various human cancers and its molecular implications as an m6A demethylase, indicated by the regulation of different targets and affected signaling pathways. (A) FTO upregulation in AML modulates ASB2/RARA levels, hence, promoting AML cellular viability and proliferation. FTO inhibition via FB23-2 inhibits AML progression, by upregulating ASB2/RARA and downregulating MYC/CEBPA. (B) Given FTO’s tumorigenic role in glioblastoma, its inhibition via MA2 suppresses GSCs self-renewal and tumorigenesis. Additionally, inhibiting FTO using R-2HG exerts anti-tumor functions in lethal glioma and leukemia by modulating MYC/CEBPA levels. (C) FTO upregulation in CSCC induces chemo-radiotherapy resistance by increasing β-catenin and activating the ERCC1 pathway. E2F1/MYC are also upregulated targets in CSCC that promote oncogenic functions. (D) FTO overexpression in lung cancer upregulates MZF1 and USP7 levels via modulating their m6A levels. Thus, promoting lung cancer cells’ proliferation, invasion, and colony formation, while inhibiting their apoptosis. (E) In breast cancer cells, FTO upregulation demethylates BNIP3 as a downstream target and downregulates it in a YTHDF2-independent manner, and promoting oncogenic roles like metastasis. (F) In melanoma, upregulated FTO and reduced m6A levels, stabilize melanoma-promoting genes namely, PDCD1, CXCR4, and SOX10 in a YTHDF2-mediated process. Consequently, increasing melanoma migration, proliferation, and immunotherapy resistance. (G) In endometrial carcinoma, higher HOXB13 expression activates Wnt signaling and promotes metastasis. (H) In pancreatic cancer, FTO overexpression leads to upregulating c-MYC oncogene and enhancing pancreatic cancer proliferation. FTO’s additional role as a tumor suppressor in pancreatic cancer is discussed in the text. Abbreviations: ASB2: Ankyrin repeat and SOCS box protein 2; RARA: retinoic acid receptor α; GSCs: glioblastoma stem cells; MA: Meclofenamic acid; R-2HG: R-2-hydroxyglutarate; ERCC1: excision repair cross-complementation group 1; MZF1: Myeloid Zinc Finger Protein 1; USP7: ubiquitin-specific protease-7; BNIP3: BCL2 Interacting Protein 3; HOXB13: homeobox transcription factor.
Ijms 23 03800 g002
Table 1. Summary of studies on FTO molecular mechanisms in relation to adipogenesis.
Table 1. Summary of studies on FTO molecular mechanisms in relation to adipogenesis.
FindingYearFTO ManipulationCell Model UsedReference
FTO regulates ghrelin m6A levels and subsequently mRNA abundance2013N/A
FTO SNP rs9939609 (risk allele A)
Peripheral blood cells from AA subjects used in m6A methylation investigationKarra et al.
[52]
FTO deficiency was reported to upregulate uncoupling protein 1 (Ucp-1) and subsequently enhanced mitochondrial uncoupling and energy expenditure; resulting in induction of brown adipocyte phenotype2013FTO knockdown via short hairpin RNA (shRNA)Human adipocytes and pre-adipocytes (from 3 healthy donors)Tews et al.
[59]
Inverse correlation during adipogenesis between m6A levels and FTO gene expression; the regulatory role of FTO in splicing of RUNX1T12014FTO depletion via siRNA3T3-L1 pre-adipocytesZhao et al.
[10]
Negative regulation of lipid accumulation by m6A levels in porcine adipocytes2015FTO knockdown with shRNAPorcine adipocytesWang et al.
[49]
FTO adipogenic effect is mediated through mitotic clonal expansion (MCE); an early stage in adipogenesis2015FTO knockdown via siRNAPrimary adipocytes and MEFs from genetically modified mice modelsMerkestein et al.
[46]
FTO functional requirement for pre-adipocyte differentiation2015FTO knockdown via siRNA3T3-L1 pre-adipocytesZhang et al.
[54]
FTO SNP rs1421085 results in an increase in energy-storing white adipocytes, reduced mitochondrial thermogenesis, and increased lipid storage2015FTO SNP rs1421085 (T-to-C)/causality further investigated using CRISPR/Cas9 genome editingHuman adipocyte progenitor cellsClaussnitzer et al.
[60]
Novel regulatory mechanism of FTO as m6A demethylase in regulating lipid accumulation in skeletal muscle cells2017FTO knockdown via siRNAMouse myoblast cell-line C2C12; Wild-type and obese mice models used for muscle collection and analysisWu et al.
[61]
Epigallocatechin gallate (ECGC) shown to inhibit adipogenesis by inhibiting the MCE stage through targeting FTO as an m6A demethylase and in an m6A-YTHDF2-dependent manner2018FTO reduced via:
  • an inhibitor (ECGC)
  • siRNA
3T3-L1 pre-adipocytesWu et al.
[62]
Zinc Finger Protein 217 (Zfp217) reported as a regulator of adipogenesis by activating FTO m6A demethylase at the transcriptional level, and by interacting with YTHDF2 post-transcriptionally2019N/A
Zfp217 knockdown via siRNA
3T3-L1 pre-adipocytesSong et al.
[65]
Entacapone identified as an inhibitor of FTO by mediating metabolic regulation through FOXO1 (as a direct substrate of FTO)2019
  • An inhibitor (entacapone)
  • shRNA
Diet-induced obese (DIO) mouse model, diabetic db/db mice; and Hep-G2 hepatic cells (for FTO knockdown by shRNA)Peng et al.
[73]
FTO as an m6A demethylase reported as a regulator of autophagy and adipogenesis2019FTO knockdown via siRNA3T3-L1 pre-adipocytes; porcine primary adipocytes and FTO-KO miceWang et al.
[71]
A link between metabolism and RNA m6A methylation was illustrated, where NADP binds FTO and enhances its activity; thus promoting m6A demethylation and adipogenesis2020FTO knockdown via siRNA3T3-L1 pre-adipocytes and FTO-KO miceWang et al.
[70]
FTO deficiency promotes thermogenesis and white-to-beige adipocyte transition via YTHDC2-mediated translation and enhanced protein expression of HIF1A2021Adipose-specific FTO knockout mice (FTOAKO)FTO knockout mice (FTOAKO) and 3T3-L1 pre-adipocytesWu et al.
[72]
Table 2. Reported biochemical inhibitors of FTO enzymatic activity.
Table 2. Reported biochemical inhibitors of FTO enzymatic activity.
InhibitorOutcome(s)Limitation(s)m6A Evaluation AssayReference
RheinInhibitory activity verified using in-silico, biochemical and cellular tests
-
In-vivo specificity remains to be explored to further evaluate efficacy
Liquid chromatography–mass spectrometry assay (LC-MS/MS)[111]
Modulation of RNA methylation
and adipogenic differentiation by Rhein demonstrated to occur in a separate manner
-
In-vivo specificity remains to be explored to further evaluate efficacy
-
In-vitro tests performed using murine pre-adipocytes only (validation using human primary adipocytes is needed)
Methylated RNA immunoprecipitation sequencing (MeRIP-seq),[117]
2OG analogsBiochemical and crystallographic tests were used to illustrate the inhibitory activity of four potent inhibitors
-
No testing of inhibitory activity using in-vitro cellular assays nor in-vivo testing.
-
Inhibitory activity verified using demethylation tests on 3-methylthymine only, hence, other FTO substrates ought to be tested.
Liquid
chromatography (LC)
[112]
Meclofenamic acid (MA)Biochemical, crystallographic, and in-vitro (HeLa cells) tests confirmed binding and inhibitory activity
-
In-vivo specificity remains to be explored to further evaluate efficacy
High-performance liquid chromatography (HPLC)-based assay[91]
Newly synthesized compound (7d: N-(3,4-Dihydroxy-5-(4-chlorophenyl)-2-furanyl)-
Ethanesulfonamide)
Computational docking to confirm binding activity, and in-vitro testing (HeLa cells) to evaluate cellular m6A levels
-
In-vivo specificity remains to be explored to further evaluate efficacy (in context of m6A levels)
-
Animal models used focused on epilepsy as per study purpose; other FTO-related disorders ought to be investigated (e.g., obesity and cancer)
LC-MS/MS[114]
New synthesized compound: (compound 12: 4-[N’-(4-Benzyl-pyridine-3-carbonyl)-hydrazino]-4-oxo-but-2-enoic acid)Crystallographic, molecular modeling, biochemical and in-vitro (HeLa cells) testing confirmed binding and inhibitory activity
-
In-vivo specificity remains to be explored to further evaluate efficacy
HPLC-based assay[115]
N-CDPCB (1a)Crystallographic, molecular modeling, biochemical and in-vitro (3T3-L1 pre-adipocytes) testing confirmed binding and inhibitory activity
-
In-vitro tests performed using murine pre-adipocytes only (validation using human primary adipocytes is needed)
-
In-vivo specificity remains to be explored to further evaluate efficacy
LC-MS/MS[116]
EntacaponeInhibitory activity of this repurposed drug was affirmed after virtual screening and in-silico testing using in-vitro (hepatic Hep-G2 cells) and in-vivo (DIO mice) assays
-
High drug dosage utilized in animal models; inability to extrapolate results from conducted mice studies directly to humans
M6A-antibody pulldown assay combined with m6A RNA seq; helped in evaluating m6A levels as well as identifying a substrate using transcriptome profiles[73]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Azzam, S.K.; Alsafar, H.; Sajini, A.A. FTO m6A Demethylase in Obesity and Cancer: Implications and Underlying Molecular Mechanisms. Int. J. Mol. Sci. 2022, 23, 3800. https://doi.org/10.3390/ijms23073800

AMA Style

Azzam SK, Alsafar H, Sajini AA. FTO m6A Demethylase in Obesity and Cancer: Implications and Underlying Molecular Mechanisms. International Journal of Molecular Sciences. 2022; 23(7):3800. https://doi.org/10.3390/ijms23073800

Chicago/Turabian Style

Azzam, Sarah Kassem, Habiba Alsafar, and Abdulrahim A. Sajini. 2022. "FTO m6A Demethylase in Obesity and Cancer: Implications and Underlying Molecular Mechanisms" International Journal of Molecular Sciences 23, no. 7: 3800. https://doi.org/10.3390/ijms23073800

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop