Next Article in Journal
The Transcriptomic Landscape of Prostate Cancer Development and Progression: An Integrative Analysis
Next Article in Special Issue
Tumor Hypoxia as a Barrier in Cancer Therapy: Why Levels Matter
Previous Article in Journal
An Analysis of Vertebral Body Growth after Proton Beam Therapy for Pediatric Cancer
Previous Article in Special Issue
Cyclic Hypoxia: An Update on Its Characteristics, Methods to Measure It and Biological Implications in Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Roles of HIF and 2-Oxoglutarate-Dependent Dioxygenases in Controlling Gene Expression in Hypoxia

1
Department of Molecular Physiology and Cell Signalling, Institute of Systems, Molecular and Integrative Biology, University of Liverpool, Liverpool L69 7ZB, UK
2
Centre for Gene Regulation and Expression, School of Life Sciences, University of Dundee, Dundee DD1 5EH, UK
*
Author to whom correspondence should be addressed.
Submission received: 24 November 2020 / Revised: 12 January 2021 / Accepted: 15 January 2021 / Published: 19 January 2021
(This article belongs to the Special Issue Novel Perspectives on Hypoxia in Cancer)

Abstract

:

Simple Summary

Hypoxia—reduction in oxygen availability—plays key roles in both physiological and pathological processes. Given the importance of oxygen for cell and organism viability, mechanisms to sense and respond to hypoxia are in place. A variety of enzymes utilise molecular oxygen, but of particular importance to oxygen sensing are the 2-oxoglutarate (2-OG) dependent dioxygenases (2-OGDs). Of these, Prolyl-hydroxylases have long been recognised to control the levels and function of Hypoxia Inducible Factor (HIF), a master transcriptional regulator in hypoxia, via their hydroxylase activity. However, recent studies are revealing that such dioxygenases are involved in almost all aspects of gene regulation, including chromatin organisation, transcription and translation.

Abstract

Hypoxia—reduction in oxygen availability—plays key roles in both physiological and pathological processes. Given the importance of oxygen for cell and organism viability, mechanisms to sense and respond to hypoxia are in place. A variety of enzymes utilise molecular oxygen, but of particular importance to oxygen sensing are the 2-oxoglutarate (2-OG) dependent dioxygenases (2-OGDs). Of these, Prolyl-hydroxylases have long been recognised to control the levels and function of Hypoxia Inducible Factor (HIF), a master transcriptional regulator in hypoxia, via their hydroxylase activity. However, recent studies are revealing that dioxygenases are involved in almost all aspects of gene regulation, including chromatin organisation, transcription and translation. We highlight the relevance of HIF and 2-OGDs in the control of gene expression in response to hypoxia and their relevance to human biology and health.

1. Introduction

The importance of oxygen for energy production in multicellular organisms has been appreciated since the identification of the mechanism of oxidative phosphorylation located in the mitochondria. Reduction in oxygen availability, or hypoxia, is therefore either a danger signal or a cue for physiological processes such as development. Every cell has a different threshold for mounting a hypoxia response, although the exact mechanisms dictating this threshold are currently not well understood. Given the importance of oxygen, cells have evolved sophisticated mechanisms to sense and respond to hypoxia, in order to minimise damage, preserve energy and, when possible, adapt to the new oxygen supply normality.
The main transcription factor activated under low oxygen conditions, called Hypoxia Inducible Factor (HIF), was identified in 1992 [1]. HIF is composed of a heterodimer of HIF-α (of which there are three isoforms encoded by three different genes, HIF-1α, HIF-2α and HIF3-α) and HIF-1β [2]. HIF3-α, unlike the other HIF-αs, lacks a transactivation domain, and is thought to act either as a dominant-negative or transcription-independent regulator of the hypoxic response [3]. HIFs control many genes, most of which are crucial for cell survival and adaptation to low oxygen conditions [2]. Under pathological conditions, such as cancer and altitude sickness, induction of some of these genes by the HIF transcription factors has been linked to disease progression and treatment resistance [4]. In addition, HIF can also be induced by non-oxygen dependent mechanisms, such as inflammation [5]. This is particularly relevant for human cancers, where hypoxia and inflammation often co-occur [6].
The detailed mechanism leading to the activation of HIF was unravelled in 2001 [7,8], building on accumulated knowledge of the role of Von Hippel–Lindau Tumor Suppressor (VHL) in this process [9,10]. HIF-α, under normal oxygen conditions, is continually transcribed and translated, but rapidly degraded by the ubiquitin dependent proteasomal system (Figure 1). Ubiquitination is promoted by the E3-ligase composed of VHL, Ring-Box 1 (RBX1), Cullin 2 (CUL2) and Elongin B/C (ELOCB/C) [4]. VHL affinity toward HIF-α is dramatically increased by the presence of a specific post-translational modification, more precisely two separate proline hydroxylation events (Figure 1), mediated by Prolyl-Hydroxylases (PHDs). PHDs are part of the 2-oxoglutarate (2-OG) dependent dioxygenase (2-OGD) superfamily of enzymes, requiring oxygen, iron (II) and 2-OG for activity [11]. Mammals possess three PHDs, PHD2 (gene name EGLN1), PHD3 (gene name EGLN3) and PHD1 (gene name EGLN2) [7].
Biochemical characterisation revealed that PHDs have low affinity for molecular oxygen. Low affinity for molecular oxygen signifies that when oxygen availability is reduced, these enzymes are quickly inhibited, leading to HIF stabilisation and activation of target genes. PHD inhibition in hypoxia has resulted in them being termed molecular oxygen sensors in the cell [12]. Additionally, many more 2-OGDs have been identified, most of which act independently of HIF but play a role in coordinating the cellular response to hypoxia. These include Factor Inhibiting HIF1 (FIH), Jumonji C (JmjC)-domain containing demethylases (JmjC demethylases) (which demethylate both histones and non-histone proteins), Ten-Eleven Translocation (TET) enzymes (mediators of DNA-demethylation), and RNA-demethylases (reported in vitro for 2-ODGs are shown is Table 1 and Supplementary Table S1). Given the function of some of these enzymes, it is conceivable that hypoxia could influence all aspects of gene expression, from chromatin structure and epigenetics to RNA biology, translation and protein turnover (Figure 2). This perfectly equips the cell when faced with hypoxia. Under such conditions, the cell must make a coordinated effort to allow for restoration of oxygen homeostasis, while reducing energy expenditure if it is to survive.
Genetic models in several model organisms have helped identify the key roles of HIFs as well as 2-OGDs in development and disease (Table 2; Supplementary Table S2). Furthermore, genome wide mapping techniques such as chromatin immunoprecipitation followed by sequencing (ChIP-seq), RNA sequencing (RNA-seq), and Chromatin capture have more recently been used to better understand how cells responds to changes in oxygen, but also in response to 2-OGD inhibition ([13,14] and reviewed in [15,16]).
In this review, we highlight the importance of oxygen sensing in coordinating an efficient response to hypoxia. We discuss the relevance of HIF transcription factors, and roles of 2-OGDs in controlling almost all aspects of gene expression from chromatin structure, to transcription, translation and post-translational modifications.

2. Effects of Hypoxia on Gene Transcription

It is now appreciated that the cellular and organism response to hypoxia involves profound changes to gene expression, with vast changes in gene transcription being detected in all systems studied.

Hypoxia-Induced Changes to Transcription Are Largely Mediated by HIF

As mentioned earlier HIFs are the master regulators of gene transcriptional changes in hypoxia and this has been extensively reviewed previously [2,11,203]. The discovery of hypoxia responsive genes and HIF targets has been driven greatly by transcript profiling and genome occupancy technologies, including microarrays, RNA-seq and ChIP-seq (reviewed in [15,16,204]). These analyses of publicly available transcriptional datasets [205,206] have shed light on cell type differences in hypoxia responsive genes and HIF targets on a genome wide scale, as well as identifying hypoxic signatures conserved across multiple cell types. Why different cell types have different transcriptional responses to hypoxia and HIF target genes is an important question for the field, with chromatin structure organisation thought to be a major factor in conferring specificity. Further to HIF isoform expression and activity, evidence points towards pre-established chromatin accessibility and local chromatin environment, including RNA pol II availability, pre-existing promoter enhancer interactions at HREs, and HRE DNA methylation status, as cell-type specificity determinants of hypoxia transcriptional responses (reviewed in [15,16,204,207]).
Whilst most HIF binding sites are at proximal promoters, binding to distal intergenic regions also occurs and HIF can regulate transcription of long genomic intervals, interacting at promoter-enhancer loops [208,209,210]. Although there is no doubt HIF is the main transcription factor controlling hypoxia-induced transcriptional changes, there is involvement of other transcription factors, including Nuclear Factor-κB (NF-κB), Tumor Protein p53 (p53), MYC Proto-Oncogene (MYC) and Activator Protein 1 (AP1), which function in the regulation of the hypoxia response via HIF-dependent and independent pathways (reviewed in [211]). In addition, HIF mostly acts as an activator of transcription, and thus most of the observed hypoxia-induced gene silencing is either independent of HIF, or via indirect mechanisms, including through the actions of chromatin remodeller complexes, co-repressor complexes or induction of other transcription factors (reviewed in [212]). HIF induced genes are involved in a variety of different pathways involved in restoration of oxygen homeostasis [2,11,203]. Importantly, many 2-OGDs, including PHD2 [213] and PHD3 [214], several JmjC demethylases (reviewed in [215]), and TET1 [216] and TET3 [216,217], are HIF target genes that have been shown to be transcriptionally upregulated in response to hypoxia.
Although we now have access to a vast number of transcriptomic studies of cells in hypoxia, very few studies have investigated the proteomic changes under such an important stress, using techniques ranging from the “old-fashioned” two-dimensional electrophoresis (2-DE) coupled with MALDI-TOF-TOF-MS [218,219,220], to the more accurate and robust quantitative multiplexed proteomics workflow [221,222,223,224], which usually combines isobaric labelling, two-dimensional liquid chromatography (2D-LC) and high resolution MS. The few studies available have shown that hypoxia exposure of different cells and tissues possesses broad effects at the whole proteome level, including changes to the protein expression of annexin family [218], glycolytic and antioxidant enzymes [219,220], transcription factors [225], heat shock proteins, S100 family proteins [221], and also other proteins involved in the TCA-cycle [226], metabolism [227] and immune response [222]. A proteomic study has revealed novel non-HIF hypoxia regulators, including the chromatin organizer protein Heterochromatin Protein 1 Binding Protein 3 (HP1BP3), which mediates chromatin condensation [223]. The use of multi-omics techniques (transcriptomics, proteomics and metabolomics) and analysis using integrated bioinformatics identified consistent changes to proteins and metabolites in heart tissues under antenatal hypoxia. These proteins and metabolites are involved in energy metabolism, oxidative stress and inflammation-related pathways, required for the reprogramming of the mitochondrion [224].

3. Chromatin Regulation in Hypoxia

Central to the hypoxia response is the activation of a dynamic transcriptional programme. HIF transcription factors are the primary mediators of hypoxia induced gene transcriptional changes (reviewed in [11]). Further to HIF stabilisation and activation under low oxygen tensions, the chromatin landscape also plays a complex role in coordinating hypoxia inducible changes to gene transcription. Most aspects of chromatin regulation are altered in response to low oxygen, including histone methylation and acetylation, DNA methylation, actions of chromatin remodeller complexes and non coding RNAs, histone eviction and incorporation of histone variants, and chromatin accessibility (reviewed in [15,16,203]). However, this is still a vastly unexplored aspect of the hypoxia response. As mentioned above, in addition to PHDs, TETs (mediators of DNA demethylation), and JmjC demethylases (histone and non-histone protein demethylases), are also 2-OGDs. Recent studies demonstrate the potential of TETs and JmjC demethylases to function as molecular oxygen sensors, directly linking oxygen sensing to transcriptional control via epigenetics in cells. Below we summarise DNA and histone methylation changes in hypoxia, with a focus on oxygen sensing mechanisms via TETs and JmjC demethylases. (Figure 3).

3.1. JmjC Demethylases and Chromatin Regulation in Hypoxia

Histone methylation is a recognised mechanism controlling chromatin structure and is associated with regulation of gene transcription, with some marks clearly leading to open chromatin, while others are firmly associated with closed conformation [228]. The family of enzymes predominantly responsible for histone demethylation are JmjC demethylases, which are part of the JmjC-domain containing group of 2-OGDs that includes demethylases and hydroxylases (JmjC 2-ODGs)). In vitro studies demonstrate the varied oxygen sensitivities of these enzymes, some are potentially direct molecular oxygen sensors (Table 1 and Supplementary Table S1). Oxygen affinities, oxygen availability and protein expression levels will likely dictate JmjC demethylase activities in hypoxia. Several groups have reported increases in total levels of histone methylation modifications in response to hypoxia across a range of human and mouse cell types and human tumours using immunoblotting, immunohistochemistry, immunofluorescence and quantitative proteomics ([229], reviewed in [16]). ChIP-sequencing approaches have revealed site-specific, hypoxia-induced changes in Histone (H)3 Lysine (K)4 trimethylation (me3) [230,231], H3K36me3 [230] and H3K27me3 [25,231], which correlate with changes in gene expression. There is now evidence, through the use of in vitro and in cell histone demethylation assays, in coordination with mutagenesis analysis, gene expression analysis and histone methylation analysis, that Lysine Demethylase (KDM) 6A (KDM6A) [25] and potentially KDM5A [230], which demethylate H3K27me3 and H3K4me3 respectively, are inhibited by reduced oxygen levels in hypoxia. These result in increased histone methylation modifications, which coordinate hypoxia inducible gene transcriptional changes (Figure 3) and hypoxia induced cellular responses. Specifically, KDM6A inhibition in hypoxia triggers hypermethylation of H3K27me3 at a subset of hypoxia repressed gene promoters, reducing their expression. Conversely, potential KDM5A inhibition in hypoxia, triggers H3K4me3 hypermethylation at a subset of hypoxia inducible gene promoters, this precedes increases in their expression in hypoxia and is required for their full transcriptional activation in hypoxia. In cell and in vitro H3K9me3 demethylation assays have also revealed that the demethylase activity of KDM4A is highly sensitive to oxygen concentrations over physiologically relevant ranges [24]. Although there are a range of reported in vitro oxygen affinities, (Supplementary Table S1) the data from cells show that KDM4A activity is sensitive to oxygen levels. There is evidence that reduced histone demethylase activity of KDM4A in severe hypoxia (<0.5% oxygen), along with KDM4C, triggers increased expression of a subset of androgen receptor target genes [232]. Thus, KDM4A can also be classed as an oxygen sensor. Interestingly, KDM4A has been shown to positively regulate HIF-1α levels via H3K9me3 demethylation at the HIF1A gene locus, this effect is observed in mild hypoxia (2% oxygen), but impaired at severe hypoxia (<0.1% oxygen) [233]. This may provide a mechanism of maintaining HIF-1α levels in conditions of mild hypoxia specifically, where there is still residual PHD activity. Future work should investigate if the oxygen sensitive H3K9me3 demethylase activity of KDM4A is linked to control of gene expression and chromatin regulation in hypoxia. Importantly, some JmjC demethylases remain active at low oxygen concentrations and function in hypoxia through their histone demethylase activity. KDM4C [234] and KDM3A [235,236] display HIF coactivator activity in hypoxia via demethylation of H3K9 at HIF target gene promoters, facilitating transcriptional activation at these genes. The co-activator role of KDM3A in hypoxia, via retained histone demethylation activity appears to be context dependent, as other groups have reported reduced activity KDM3A in hypoxia [237]. Furthermore, many JmjC histone demethylases are HIF target genes that are upregulated in hypoxia (reviewed in [215]), this is thought in part to be a compensatory mechanism to counteract reduced demethylase activity, acting as a hypoxia feedback loop similar to what is seen with transcriptional upregulation of PHD2/3 by HIF. Thus, there is complex crosstalk between histone methylation, gene expression and hypoxia, mediated in part by JmjC demethylases. However, further characterisation of the oxygen sensitives of JmjC demethylases is needed.

3.2. TET Mediated DNA Demethylation Functions in Hypoxia

TETs, of which there are three variants—Ten Eleven Ten translocation Methylcytosine Dioxygenase 1 (TET1), TET2 and TET3—are 2-ODGs that function as hydroxylases, mediating mammalian DNA demethylation through catalysing the oxidation of 5-methylcytosine (5mc) to 5-hydroxymethylcytosine (5hmC), 5-formylcytosine (5fC) and 5-carboxylcytosine (5caC). These TET oxidised derivatives of 5mc can be demethylated by mechanisms of active and passive demethylation (reviewed in [238]). DNA methylation can repress gene transcription, consequently tumour hypoxia results in aberrant DNA methylation profiles promoting tumour suppressor gene silencing (hypermethylation) [239] and oncogene activation (hypomethylation) [240]. Recently, using in vitro biochemical binding assays, in vivo studies on HIF binding and DNA methylation status in human cancer cell lines, and in silico structural modelling, D’Anna and colleagues find that DNA methylation at HREs impairs HIF binding, and HRE DNA methylation status is a key factor in determining cell type specific transcriptional responses to hypoxia [241]. There is heterogeneity regarding the effects of hypoxia, both in cell models and in tumours, on global DNA methylation levels and TET activity (reviewed in [203,242]). As such, whether TETs are impaired or functionally active in hypoxia, and the consequences this has for gene transcription appear highly context dependent. Researchers have shown TET activity in hypoxia, and HIF-dependent TET upregulation and coactivator functions have been demonstrated at hypoxia-inducible genes (reviewed in [203,242]). TET activity in low oxygen environments is supported by in vitro oxygen affinities of TET1 and TET2 (Table 1 and Supplementary Table S1), as well as the known roles of TETs in the bone marrow and during development where oxygen tensions are low [243,244]. Conversely, Thienpont et al. showed that severe hypoxia (0.5% oxygen) in human and murine cells and tumour hypoxia in multiple human tumours causes DNA hypermethylation at gene promoters correlating with gene silencing at a subset of hypoxia repressed genes and gene silencing linked to hypoxia associated tumour progression. DNA hypermethylation was attributed to oxygen dependent reduction in TET1 and TET2 activity in hypoxia, with a 50% reduction in activity observed at 0.3% oxygen for TET1 and 0.5% for TET2 in vitro. Thus, TET1 and TET2 may be characterised as tumour oxygen sensors, and depending on the context of oxygen deprivation, may remain active in hypoxia environments or display inhibition. However, more work is needed to establish the oxygen dependence of TET activity in cells and in vivo and the physiological contexts in which TETs can sense changes in oxygen availability, as well as the consequences this has for DNA methylation, gene transcription and cellular responses. Indeed, the seemingly contradictory roles for TETs in hypoxia from studies to date may be dependent on the different cell models used and timing/severity of hypoxic stimulus.
While there is growing evidence for a dynamic role of chromatin/epigenetics in sensing and responding to hypoxia to facilitate transcriptional changes, via HIF-dependent and -independent mechanisms, efforts to elucidate molecular mechanisms underpinning such changes and the extent to which chromatin/epigenetic changes are required for coordinating hypoxia/HIF transcriptional effects are ongoing. The discoveries of oxygen sensing by TETs and JmjC demethylases provide an exciting link between oxygen availability and chromatin regulation, and future work on oxygen sensing by chromatin will be essential for better understanding hypoxia driven processes.

4. Effects of Hypoxia on Protein Levels

Protein levels and function are key aspects to achieve the appropriate hypoxia response. Although transcription is important, mechanisms controlling protein levels and function supersede any changes in transcriptional output. In hypoxia, mechanisms exist that control translation, but also post-translation aspects of protein function.

4.1. Translation Is Globally Repressed in Response to Hypoxia

In addition to the regulation of gene transcription in hypoxia, gene expression is also controlled through regulation of translation (Figure 4). The cellular response to hypoxia includes a reduction in the energy expenditure of the cell due to limited ATP production through oxidative phosphorylation. This adaptation results in a reversible global decrease in energy-expensive protein synthesis (reviewed in [2]). This inhibition of translation is a highly regulated response to low oxygen levels preceding ATP depletion (reviewed in [2]).
Global inhibition of protein expression is largely regulated at the point of translation initiation through two pathways. Firstly, Mechanistic Target of Rapamycin Kinase (mTOR) is inhibited by DNA Damage Inducible Transcript 4 (DDIT4) (a HIF target gene). DDIT4-dependent release of TSC Complex Subunit 2 (TSC2) from 14-3-3 binding proteins leads to mTOR inhibition. This allows the formation of an active TSC1-TSC2 dimer that inhibits the phosphorylation of Ribosomal Protein S6 Kinase B1 (RPS6KB1) by the mTOR protein complex, which in turn inhibits the phosphorylation of Ribosomal Protein S6 (RPS6), part of the 40S ribosomal subunit required for initiation [245]. mTOR inhibition also causes hypophosphorylation of Eukaryotic Translation Initiation Factor (eIF) 4E Binding Protein 1 (eIF4EBP1), allowing the sequestering of eIF4E and decrease of 5’cap-dependent initiation [246,247]. Secondly, PKR-like Endoplasmic Reticulum Kinase (PERK) (gene name EIF2AK3) is phosphorylated and activated, which subsequently phosphorylates eIF2α at S51, causing effective inactivation [248]. Phospho-eIF2α prevents binding with eIF2B for exchange of GDP for GTP, therefore remaining in an inactive state and preventing subsequent rounds of translation from the mRNA [249]. eIF2α normally recruits the initiator aminoacylated tRNA to the 40S ribosome, thus limiting global initiation of translation. This second mechanism is independent of HIF and as of yet has not been linked to any 2-OGD.
Inhibition of translation is also regulated at the stage of polypeptide elongation. Elongation is inhibited by phosphorylation of Eukaryotic Elongation Factor 2 (eEF2) at T56 by eEF2 Kinase (eEF2K) [250]. This process has been shown to be dependent on mTOR and 5′-AMP-activated protein kinase catalytic subunit alpha-1 (AMPK) (gene name PRKAA1) [251,252]. Interestingly, eEF2 kinase (eEF2K) is also regulated by hydroxylation by PHD2 at P98 in an oxygen-dependent manner [253]. In hypoxia, when PHD2 inhibited, eEF2K activity is induced.
In addition to PHD2 dependent hydroxylation, there are several other hydroxylation reactions involved in the regulation of translation, catalysed by other 2-OGDs. Hydroxylation is important for the biosynthesis of tRNAPhe with position 37 requiring a hypermodified nucleoside Wybutosine (yW) that can be hydroxylated to form hydroxywybutosine (OHyW) by the JmjC hydroxylase, TRNA-YW Synthesizing Protein 5 (TYW5), which maintains translational fidelity. It is currently unknown whether TYW5 is responsive to the level of oxygen, but its transcription is decreased in hypoxia, [254] linking hypoxia to a decreased accuracy of translation, which globally decreases the successful translation of proteins [255].
The rate and accuracy of translation are positively regulated by hydroxylation of the central translation machinery. JmjC hydroxylases Ribosomal Oxygenase 2 (RIOX2) and RIOX1 hydroxylate histidyl residues in ribosomal proteins, with RIOX2 and RIOX1 hydroxylating Ribosomal Protein L (RPL) 27a (RPL27A) and (RPL8), respectively. The hydroxylation occurs at residues close to the peptidyl transfer centre, thereby increasing translation efficiency [256]. RIOX1 and RIOX2 transcription is reduced in hypoxia [254,256]. Furthermore, RPL8 hydroxylation is also reduced in hypoxia [256]. However, it is not yet known whether these enzymes are inhibited by low oxygen levels, or lower hydroxylation is solely due to lower transcription. Additionally, hydroxylation of 40S Ribosomal Protein S23 (RPS23) by the 2-OGD, 2-Oxoglutarate and Iron Dependent Oxygenase Domain Containing 1 (OGFOD1), is required for efficient translation [257]. OGFOD1 transcription is also decreased in hypoxia, but the enzyme remains mostly active, even in acute hypoxia [258], suggesting this mechanism is not through direct 2-OGD oxygen sensing.
Efficient decoding of the mRNA during translation requires the JmjC hydroxylase, AlkB Homolog 8, TRNA Methyltransferase, ALKBH8, which hydroxylates tRNA at the wobble position [259,260]. This 2-OGD has yet to be linked to hypoxia, though it would be interesting to investigate its oxygen sensitivity. Finally, lysyl hydroxylation of eukaryotic release factor 1 (eRF1) by the JmjC hydroxylase Jumonji Domain Containing 4 (JMJD4) is required for proper termination of translation [261], although its activity is not significantly inhibited in hypoxia.

4.2. Utilising Proteomics for the Identification of Non-Histone Protein PTMs

Proteomics approaches have revealed hypoxia induces changes to many post translational modifications (PTMs) on non-histone proteins, such as proline hydroxylation [262,263] (regulating protein levels and interactions), phosphorylation [264,265], SUMOylation [266], acetylation [267], glycosylation [268], nitration [269] and nitrosylation [270], all of which regulate protein functions in different ways.
As one of the most widely studied PTMs, phosphorylation of some transcriptional factors and regulators has been found to be changed under various hypoxic conditions, including CAMP Responsive Element Binding Protein 1 (CREB1), NFKB Inhibitor Alpha (NFKBIA), a regulator of NF-κB, and HIF (reviewed in [271]). More recently, through the analysis of phospho-proteomics in renal clear cell carcinoma cells under a VHL-independent hypoxic response, up-regulation of known biomarkers of RCC and signalling adaptor were found. Meanwhile, such hypoxic responses decreased the phosphorylation on intracellular Carbonic Anhydrase 2 (CA2), which might be an unusual way to control the CA2 expression and enhance the activity of the NF-κB pathway, resulting in loss of VHL [264].
In recent years, non-HIF targets have been identified to be hydroxylated on prolines by PHDs (reviewed in [15]), resulting in their degradation and/or changes to downstream activity including Centrosomal Protein 192 (CEP192) [272], Forkhead Box O3 (FOXO3) [273] and Adenylosuccinate Lyase (ADSL) [274] by PHD1, Actin Beta (ACTB) by PHD3 [275], and AKT Serine/Threonine Kinase 1 (AKT1) [93], TANK Binding Kinase 1 (TBK1) [276] and Scm-like with Four Malignant Brain Tumour Domains 1 (SFMBT1) [277] by PHD2 [278]. Despite these exciting studies, a recent study investigating around 20 of the reported non-HIF substrates, failed to detect any hydroxylation in vitro [279]. This adds to the complexity of identifying novel PHD substrates. Interestingly, a new study indicated that prolyl-hydroxylation could be crucial for GMGC kinase activation [280]. This could imply an intricate interplay between these two types of PTMs, suggesting yet another role for oxygen-dependent signalling in the cell. Thus, unbiased proteomic studies on novel PTMs sites [262,281], system-wide analysis of PHDs substrates other than HIF-α [263] and crosstalk of PTMs on PHDs targets in response to hypoxia are now emerging.

5. Other Potential Roles of 2-OGDs in the Hypoxia Response

Further to known and potential oxygen sensing roles of JmjC 2-ODGs (JmjC demethylases and hydroxylases) in the regulation of chromatin and translation discussed earlier, there are other functions of JmjC 2-ODGs that may influence the hypoxia response (Figure 5). One of the most prominent of such enzymes is the JmjC 2-OGD, JMJD6, Arginine Demethylase and Lysine Hydroxylase, which has unique activities as both an arginine demethylase (acting on mono- and symmetric/asymmetric di-methylated arginines) and lysine hydroxylase [282,283]. JMJD6 expression is increased in hypoxic conditions in the placenta and can downregulate HIF-1α [284], though it has been found to operate in diverse pathways. JMJD6 can promote the formation of stress granules through demethylation of Arg-435, Arg-447 and Arg-460 of G3BP Stress Granule Assembly Factor 1 G3BP1, and subsequent de-repression of the protein, resulting in the cytoplasmic sequestering of stalled mRNA-ribosome complexes to reversibly prevent mRNA degradation [285,286]. This would allow a fast re-start of protein synthesis when oxygen homeostasis is restored. JMJD6 also regulates mRNA splicing through hydroxylating the splicing regulatory (SR) proteins LUC7 Like 2, Pre-MRNA Splicing Factor (LUC7L2) on Lys-269, U2 Small Nuclear RNA Auxiliary Factor 65 (U2AF65) on Lys-15 and Lys-276 [287], and multiple fragments of Serine and Arginine Rich Splicing Factor 11 (SRSF11) between residue 271 to 372 [283]. The SR proteins are involved in exon definition and alternative splicing, with SRSF11 hydroxylation resulting in skipping of most 5’ exon, and hydroxylation of U2AF65 possibly enacting pre-mRNA looping in order to present to the splicing machinery different cis splice enhancer or silencer sequences [288]. However, this only occurs for selected mRNAs and is not a global effect [288]. Nevertheless, this mechanism would allow selection of alternate splice variants as a response to hypoxia. JMJD6 can also interact with both Bromodomain Containing 4 (BRD4) and the positive Transcription Factor Elongation Factor b (P-TEFb) complex [289], eventually resulting in the release of paused DNA polymerase II and resumption of mRNA synthesis at specifically regulated genes [288]. This implies that hypoxia could use this mechanism to stall transcription of genes that are not required for the stress response to hypoxia and would allow a re-start of gene expression when oxygen levels are restored. JMJD6 hydroxylates p53 on Lys382, repressing its transcriptional activity [290]. JMJD6 also auto-hydroxylates itself on the N-terminus and this is required for homo-oligomerisation [291]. The function of the homomultimer of JMJD6 is unknown, but speculated to play a role as a scaffolding protein in the nucleus [292].
Another JmjC hydroxylase, KDM8, can hydroxylate arginine residues in both RCC1 Domain Containing 1 (RCCD1) and RPS6 [293]. KDM8 also has amine- and endo-peptidase activity with specificity towards histone H3 monomethyl-lysine residues, especially of the histone variant H3F3A [294], and arginine methylated histone tails [295,296]. It had previously been reported that KDM8 can demethylate H3K36me2 [297], though considering its arginine demethylation is achieved through trimming of the histone tail, this may be the mechanism of observed decreases in lysine methylation. Although not necessarily dependent on its hydroxylation activity, KDM8 is required for cell proliferation and chromosomal stability [298], and can negatively regulate p53, affecting gene expression and control cell cycle and proliferation [299,300]. Recently, a biochemical function has been assigned to JMJD7 as a lysyl hydroxylase, which targets Developmentally Regulated GTP Binding Protein 1 (DRG1) and DRG2 that are part of the Translation Factor (TRAFAC) family of GTPases, and could affect their binding with messenger, or ribosomal RNA, though this requires further investigation [301]. Similar to KDM8, JMJD7 also possesses endo- and amine-peptidase activity, preferentially cleaving mono- and di-methylated arginine residues of H2, H3, and H4 [295,296].
Several JmjC demethylases have been found to have additional, non-histone substrates. For example, KDM2A represses NF-κB activity via demethylation of RELA, providing a possible link to hypoxia and inflammation crosstalk. [302]. KDM3A was recently found to function as a molecular oxygen sensor, with its demethylation of K224me1 on peroxisome proliferator-activated receptor gamma coactivator-1 alpha (PGC-1α) (which facilitates its activity) impaired under low oxygen conditions [237]. PGC-1α is a transcriptional coactivator promoting mitochondrial biogenesis and oxidative metabolism, the aforementioned work neatly links mitochondrial function to the hypoxia response via the oxygen sensing function of KDM3A controlling PGC-1α activity, further highlighting the importance of JmjC 2-OGDs in the hypoxia response. KDM3A can also demethylate K372me1 of p53, which decreases its association with chromatin and represses its proapoptotic activity [303]. KDM3B, in addition to its substrate H3K9me2, can also demethylate mono- or di-methylated arginine residues, H4R3me2 [127]. It is more than likely that other JmjC demethylases interact and directly demethylate additional transcription factors which may coordinate transcriptional responses to hypoxia. However, unbiased analysis is required to fully assess this aspect of hypoxia-induced gene regulation. The KDM4 family of demethylases appears to have higher activity against trimethylated non-histone targets than its canonical substrates, H3K9me2 and H3K36me3 [304]. These substrates include widely interspaced zinc finger motifs protein (WIZ), chromodomain Y-like protein (CDYL1), Cockayne syndrome group B protein (CSB, four peptides) and Histone-lysine N-methyltransferase (EHMT2). These substrates of KDM4 family are all members of transcription repression complexes, suggesting an additional role of transcriptional regulation by these demethylases. Using in vitro assays, KDM3A, KDM4E, KDM5C, and KDM6B have been shown to possess arginine demethylase activity, in addition to their lysine demethylase activity on peptide substrates [305]. Non-specific hydroxylase activities are not unusual for JmjC enzymes, so any such reported activity should demonstrate specificity and biological relevance, though it is possible that it will emerge that many lysine demethylases also possess relevant arginine demethylase activity.
Finally, a potentially new and exciting mechanism involves an epigenetic mark in RNA, methylation of 6 adenosine (m6A). mRNA m6A modification at the 5’UTR can recruit m6A binding proteins such as YTH N6-Methyladenosine RNA Binding Protein 2 (YTHDF2) in heat shock stress, which upregulates translation through binding eIF3 and the 40S subunit [306], and such a mechanism may prove true in other stress responses, such as hypoxia. The RNA demethylating enzymes FTO Alpha-Ketoglutarate Dependent Dioxygenase (FTO) and AlkB Homolog 5, RNA Demethylase (ALKBH5), are 2-OGDs, and although their ability to sense oxygen has yet to be investigated, their inhibition in hypoxia could result in an increase in global RNA methylation, which could regulate translation and RNA fate in hypoxia.

6. Relevance to Human Biology and Health

Although we currently do not know the importance of all of the 2-OGDs present in the genome, several of the key players in the hypoxia response have important functions and relevance to human biology and health. This is exemplified by the phenotypes observed in null mice, or by the presence of disease associated mutations in humans.

6.1. PHD/HIF/VHL Axis

Given the cellular functions mentioned above, it is no surprise that genetic mutations of most dioxygenases and HIFs have been implicated with human diseases. Mutations in HIF-2α and PHD2 have been found in patients with vascular pathologies, such as erythrocytosis, polycythemia, and pheochromocytoma (Table 2). As HIF mediates hypoxia adaptation responses, including the regulation of erythropoiesis and vasculogenesis, it is not surprising that mutations within the PHD/HIF/VHL axis are associated with vascular pathologies. The crucial role of HIFs in vascular pathologies is strongly demonstrated by genetic studies of mice, as highlighted in Supplementary Table S2. Knockout mice of HIF-1α or HIF-2α are embryonic lethal with vascular defects (Table 2, Supplementary Table S2), whereas the deletion of PHD2 that activates HIF signalling results in embryonic lethality in mice due to placental and heart defects (Table 2, Supplementary Table S2). VHL mutations also result in highly vascularized tumours, including pheochromocytomas, renal cell cancer carcinoma, retinal and central nervous system hemangioblastomas (Table 2). Hundreds of VHL mutations have been identified in VHL syndrome patients (listed in the Human Gene Mutation Database [307]). The homozygous VHL mutation R200W, which prevents efficient HIF-α degradation in normoxia, is found in all individuals with Chuvash Polycythemia (CP) [308]. CP is characterized by congenital erythrocytosis, and patients have been associated with pulmonary hypertension, thrombosis, vertebral hemangiomas, cerebral vascular events and other vascular abnormalities [309,310], displaying the role of VHL in HIF-dependent regulation of vasculogenesis and erythropoiesis.

6.2. 2-OGDs—Hydroxylases

Similar to PHDs, prolyl-4-hydroxylases P4HA1 and P4HA2 are hypoxia-inducible, but P4HA1/2 prolyl hydroxylation is required for different processes, that is collagen fiber formation. Consistent with their roles in collagen synthesis, P4HA1 and P4HA2 mutations are found in patients with collagen-related extracellular matrix disorders (Table 2; Supplementary Table S2). Furthermore, homozygous deletion of P4HA1 is embryonic lethal with base membrane rupture due to defective collagen IV assembly (Table 2). PAHX is another hydroxylase, but of phytanoyl-CoA; essential for breaking down phytanic fatty acid. Mutations in PAHX is well associated with Refsum disease, a rare inherited neurological disorder caused by neurotoxic phytanic acid as these mutations result in an enzymatically inactive protein, thus leading to phytanic acid accumulation.
The roles of TET1–3 in development are demonstrated in knockout mouse models (reviewed in [311]). TET1-null mice present several defects but these depend on the mode of genetic deletion. TET3 deletion results in neonatal lethality, highlighting TET3 role in development. TET3 mutations have been found in patients with intellectual disability and/or delayed global development (Table 2; Supplementary Table S2). Although somatic alterations of TET2 have been found in several cancers, these mutations are majorly associated with myelodysplastic syndromes (Table 2; Supplementary Table S2). In addition to the listed mutations in Supplementary Table S2, a study reported TET2 somatic mutations in 46 patients with myelodysplastic syndromes, myeloproliferative disorder, secondary acute myeloid leukemia, or chronic myelomonocytic leukemia [312]. Most of these mutations are predicted to lead to partial or total loss of function due to protein truncation.

6.3. 2-OGDs—JmjC Demethylases

Many of the JmjC demethylase genes have been associated with human diseases. In particular, several of them are mutated in patients with neurodevelopmental disorders, midline defects and cancers (Table 2; Supplementary Table S2). Although KDM3A is found to be mutated in infertile males [125], its role in infertility is not clear. Mutations in KDM3B are frequently implicated with intellectual disability, but also found in cancers including myeloid leukemias (Table 2). Similarly, JMJD1C mutations have been identified in individuals with autism spectrum disorder and intellectual disability. JMJD1C is also associated with congenital heart disease manifestation in 22q11.2 deletion syndrome patients. Amongst KDM4s, there are only two reports—single nucleotide substitutions of KDM4C in upper aerodigestive tract cancer and age of menarche. Mutations of KDM5B, KDM5C and KDM6B have been associated with neurodevelopment and a global developmental delay (Table 2; Supplementary Table S2). In particular, KDM5C is well recognised as an X-linked intellectual disability gene that is highly expressed in neural tissue. Mutations in PHD Finger Protein 8 (PHF8) are also associated with X-linked mental retardation and often accompanied by cleft lip/palate or autism. The phenotypes of KDM5C or PHF8 mutations in humans are reflected by the deletion of these genes in mice (Table 2). On the other hand, KDM6A mutations are frequently found in individuals with Kabuki syndrome (KS), a genetic disease with developmental delay and congenital anomalies, (Table 2) highlighting the role of KDM6A in development. In addition to mutations listed in Supplementary Table S2, others have reported gross deletions, gross duplications, or chromosomal rearrangement in patients with KS or KS-like clinical manifestations [176,178,313,314,315]. However, whether the phenotypes observed are due to loss of demethylase activity solely is currently unknown.
Oxygen regulation is pivotal for embryogenesis as most developmental processes during early embryo development occurs in a hypoxic environment [316]. Given the fact that 2-OGDs sense and respond to oxygen levels, it is not surprising that mutations affecting the activity of several of these enzymes are detrimental to embryo development, contributing to neonatal defects. The hypoxic environment is also a common feature of tumours due to aberrant vascularisation and an inadequate supply of blood [317]. Mutations or deletions of some of these 2-OGDs are implicated in cancers. The molecular mechanisms through which 2-OGDs promote cancers is not completely understood. However, altered metabolism in cancer has been linked to alteration of 2-OGDs function (reviewed in [318]). 2-OGDs can be inhibited by metabolites that are accumulated in tumours, including succinate, fumarate and D/L-2-hydroxyglutarate [318,319]. It is thus very likely that the phenotypes observed in deletions and mutations are the result of both alteration of the hypoxia response and changes in metabolism. Overall, the presence and connections of HIFs or dioxygenase mutations in human disorders and the knockout studies demonstrate the essential roles of these genes.

7. Conclusions and Future Perspectives

As our understanding of the cellular response to hypoxia advances, new aspects continue to unravel. The role of oxygen has surfaced as far broader than just an acceptor molecule in oxidative phosphorylation in the mitochondria. Through acting as a co-factor for diverse and functionally important enzymes, oxygen is mechanistically identified as a potent signalling molecule in cells. The emerging focus of the field includes new aspects of chromatin regulation, RNA biology and broad regulation of protein post-translational modifications directly controlled by oxygen levels. This advanced understanding in conjunction with development of novel therapeutic chemicals targeting dioxygenases should provide not only exciting new biological insights, but also better treatments for patients suffering from a range of diseases. One area of technological advancement that will greatly progress the field is the adaptation of novel and unbiased quantitative techniques for measuring chromatin structure, transcriptional output, proteomic changes and cellular behavior. These approaches may provide resolution to some of the persisting major questions pertaining mechanisms controlling gene expression in response to hypoxia.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2072-6694/13/2/350/s1, Supplementary Table S1: 2-OGDs reported affinities for oxygen from in vitro assays; Supplementary Table S2: Known mutations in the hypoxia signalling system in humans, and Supplementary Table references.

Funding

This work was supported by the Wellcome Trust (206293/Z/17/Z) and the University of Liverpool.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Semenza, G.L.; Wang, G.L. A nuclear factor induced by hypoxia via de novo protein synthesis binds to the human erythropoietin gene enhancer at a site required for transcriptional activation. Mol. Cell. Biol. 1992, 12, 5447–5454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Rocha, S. Gene regulation under low oxygen: Holding your breath for transcription. Trends Biochem. Sci. 2007, 32, 389–397. [Google Scholar] [CrossRef] [PubMed]
  3. Duan, C. Hypoxia-inducible factor 3 biology: Complexities and emerging themes. Am. J. Physiol. Cell Physiol. 2016, 310, C260–C269. [Google Scholar] [CrossRef] [Green Version]
  4. Kliewe, F.; Engelhardt, M.; Aref, R.; Schuller, H.J. Promoter recruitment of corepressors Sin3 and Cyc8 by activator proteins of the yeast Saccharomyces cerevisiae. Curr. Genet. 2017, 63, 739–750. [Google Scholar] [CrossRef] [PubMed]
  5. D’Ignazio, L.; Batie, M.; Rocha, S. Hypoxia and Inflammation in Cancer, Focus on HIF and NF-kappaB. Biomedicines 2017, 5, 21. [Google Scholar] [CrossRef] [Green Version]
  6. Biddlestone, J.; Bandarra, D.; Rocha, S. The role of hypoxia in inflammatory disease (review). Int. J. Mol. Med. 2015, 35, 859–869. [Google Scholar] [CrossRef] [Green Version]
  7. Epstein, A.C.; Gleadle, J.M.; McNeill, L.A.; Hewitson, K.S.; O’Rourke, J.; Mole, D.R.; Mukherji, M.; Metzen, E.; Wilson, M.I.; Dhanda, A.; et al. C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell 2001, 107, 43–54. [Google Scholar] [CrossRef] [Green Version]
  8. Ivan, M.; Kondo, K.; Yang, H.; Kim, W.; Valiando, J.; Ohh, M.; Salic, A.; Asara, J.M.; Lane, W.S.; Kaelin, W.G., Jr. HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: Implications for O2 sensing. Science 2001, 292, 464–468. [Google Scholar] [CrossRef] [PubMed]
  9. Maxwell, P.H.; Wiesener, M.S.; Chang, G.W.; Clifford, S.C.; Vaux, E.C.; Cockman, M.E.; Wykoff, C.C.; Pugh, C.W.; Maher, E.R.; Ratcliffe, P.J. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature 1999, 399, 271–275. [Google Scholar] [CrossRef]
  10. Ohh, M.; Park, C.W.; Ivan, M.; Hoffman, M.A.; Kim, T.Y.; Huang, L.E.; Pavletich, N.; Chau, V.; Kaelin, W.G. Ubiquitination of hypoxia-inducible factor requires direct binding to the beta-domain of the von Hippel-Lindau protein. Nat. Cell Biol. 2000, 2, 423–427. [Google Scholar] [CrossRef]
  11. Kaelin, W.G., Jr.; Ratcliffe, P.J. Oxygen sensing by metazoans: The central role of the HIF hydroxylase pathway. Mol. Cell 2008, 30, 393–402. [Google Scholar] [CrossRef] [PubMed]
  12. Ehrismann, D.; Flashman, E.; Genn, D.N.; Mathioudakis, N.; Hewitson, K.S.; Ratcliffe, P.J.; Schofield, C.J. Studies on the activity of the hypoxia-inducible-factor hydroxylases using an oxygen consumption assay. Biochem. J. 2007, 401, 227–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Chowdhury, R.; Candela-Lena, J.I.; Chan, M.C.; Greenald, D.J.; Yeoh, K.K.; Tian, Y.M.; McDonough, M.A.; Tumber, A.; Rose, N.R.; Conejo-Garcia, A.; et al. Selective small molecule probes for the hypoxia inducible factor (HIF) prolyl hydroxylases. ACS Chem. Biol. 2013, 8, 1488–1496. [Google Scholar] [CrossRef]
  14. Frost, J.; Ciulli, A.; Rocha, S. RNA-seq analysis of PHD and VHL inhibitors reveals differences and similarities to the hypoxia response. Wellcome Open Res. 2019, 4, 17. [Google Scholar] [CrossRef] [PubMed]
  15. Wilson, J.W.; Shakir, D.; Batie, M.; Frost, M.; Rocha, S. Oxygen-sensing mechanisms in cells. FEBS J. 2020. [Google Scholar] [CrossRef]
  16. Batie, M.; Rocha, S. Gene transcription and chromatin regulation in hypoxia. Biochem. Soc. Trans. 2020. [Google Scholar] [CrossRef]
  17. Hirsila, M.; Koivunen, P.; Gunzler, V.; Kivirikko, K.I.; Myllyharju, J. Characterization of the human prolyl 4-hydroxylases that modify the hypoxia-inducible factor. J. Biol. Chem. 2003, 278, 30772–30780. [Google Scholar] [CrossRef] [Green Version]
  18. Tarhonskaya, H.; Chowdhury, R.; Leung, I.K.; Loik, N.D.; McCullagh, J.S.; Claridge, T.D.; Schofield, C.J.; Flashman, E. Investigating the contribution of the active site environment to the slow reaction of hypoxia-inducible factor prolyl hydroxylase domain 2 with oxygen. Biochem. J. 2014, 463, 363–372. [Google Scholar] [CrossRef]
  19. Dao, J.H.; Kurzeja, R.J.; Morachis, J.M.; Veith, H.; Lewis, J.; Yu, V.; Tegley, C.M.; Tagari, P. Kinetic characterization and identification of a novel inhibitor of hypoxia-inducible factor prolyl hydroxylase 2 using a time-resolved fluorescence resonance energy transfer-based assay technology. Anal. Biochem. 2009, 384, 213–223. [Google Scholar] [CrossRef]
  20. Koivunen, P.; Hirsila, M.; Gunzler, V.; Kivirikko, K.I.; Myllyharju, J. Catalytic properties of the asparaginyl hydroxylase (FIH) in the oxygen sensing pathway are distinct from those of its prolyl 4-hydroxylases. J. Biol. Chem. 2004, 279, 9899–9904. [Google Scholar] [CrossRef] [Green Version]
  21. Tarhonskaya, H.; Hardy, A.P.; Howe, E.A.; Loik, N.D.; Kramer, H.B.; McCullagh, J.S.; Schofield, C.J.; Flashman, E. Kinetic Investigations of the Role of Factor Inhibiting Hypoxia-inducible Factor (FIH) as an Oxygen Sensor. J. Biol. Chem. 2015, 290, 19726–19742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Laukka, T.; Mariani, C.J.; Ihantola, T.; Cao, J.Z.; Hokkanen, J.; Kaelin, W.G., Jr.; Godley, L.A.; Koivunen, P. Fumarate and Succinate Regulate Expression of Hypoxia-inducible Genes via TET Enzymes. J. Biol. Chem. 2016, 291, 4256–4265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Cascella, B.; Mirica, L.M. Kinetic analysis of iron-dependent histone demethylases: Alpha-ketoglutarate substrate inhibition and potential relevance to the regulation of histone demethylation in cancer cells. Biochemistry 2012, 51, 8699–8701. [Google Scholar] [CrossRef] [PubMed]
  24. Hancock, R.L.; Masson, N.; Dunne, K.; Flashman, E.; Kawamura, A. The Activity of JmjC Histone Lysine Demethylase KDM4A is Highly Sensitive to Oxygen Concentrations. ACS Chem. Biol. 2017, 12, 1011–1019. [Google Scholar] [CrossRef] [PubMed]
  25. Chakraborty, A.A.; Laukka, T.; Myllykoski, M.; Ringel, A.E.; Booker, M.A.; Tolstorukov, M.Y.; Meng, Y.J.; Meier, S.R.; Jennings, R.B.; Creech, A.L.; et al. Histone demethylase KDM6A directly senses oxygen to control chromatin and cell fate. Science 2019, 363, 1217–1222. [Google Scholar] [CrossRef] [Green Version]
  26. Masson, N.; Keeley, T.P.; Giuntoli, B.; White, M.D.; Puerta, M.L.; Perata, P.; Hopkinson, R.J.; Flashman, E.; Licausi, F.; Ratcliffe, P.J. Conserved N-terminal cysteine dioxygenases transduce responses to hypoxia in animals and plants. Science 2019, 365, 65–69. [Google Scholar] [CrossRef]
  27. Iyer, N.V.; Kotch, L.E.; Agani, F.; Leung, S.W.; Laughner, E.; Wenger, R.H.; Gassmann, M.; Gearhart, J.D.; Lawler, A.M.; Yu, A.Y.; et al. Cellular and developmental control of O2 homeostasis by hypoxia-inducible factor 1 alpha. Genes Dev. 1998, 12, 149–162. [Google Scholar] [CrossRef] [Green Version]
  28. Ryan, H.E.; Lo, J.; Johnson, R.S. HIF-1 alpha is required for solid tumor formation and embryonic vascularization. EMBO J. 1998, 17, 3005–3015. [Google Scholar] [CrossRef] [Green Version]
  29. Compernolle, V.; Brusselmans, K.; Franco, D.; Moorman, A.; Dewerchin, M.; Collen, D.; Carmeliet, P. Cardia bifida, defective heart development and abnormal neural crest migration in embryos lacking hypoxia-inducible factor-1alpha. Cardiovasc. Res. 2003, 60, 569–579. [Google Scholar] [CrossRef]
  30. Gulsuner, S.; Walsh, T.; Watts, A.C.; Lee, M.K.; Thornton, A.M.; Casadei, S.; Rippey, C.; Shahin, H.; Consortium on the Genetics of Schizophrenia; Paartners Study Group; et al. Spatial and temporal mapping of de novo mutations in schizophrenia to a fetal prefrontal cortical network. Cell 2013, 154, 518–529. [Google Scholar] [CrossRef] [Green Version]
  31. Prior, S.J.; Hagberg, J.M.; Phares, D.A.; Brown, M.D.; Fairfull, L.; Ferrell, R.E.; Roth, S.M. Sequence variation in hypoxia-inducible factor 1alpha (HIF1A): Association with maximal oxygen consumption. Physiol. Genom. 2003, 15, 20–26. [Google Scholar] [CrossRef] [Green Version]
  32. Ollerenshaw, M.; Page, T.; Hammonds, J.; Demaine, A. Polymorphisms in the hypoxia inducible factor-1alpha gene (HIF1A) are associated with the renal cell carcinoma phenotype. Cancer Genet. Cytogenet. 2004, 153, 122–126. [Google Scholar] [CrossRef] [PubMed]
  33. Tian, H.; Hammer, R.E.; Matsumoto, A.M.; Russell, D.W.; McKnight, S.L. The hypoxia-responsive transcription factor EPAS1 is essential for catecholamine homeostasis and protection against heart failure during embryonic development. Genes Dev. 1998, 12, 3320–3324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Peng, J.; Zhang, L.; Drysdale, L.; Fong, G.H. The transcription factor EPAS-1/hypoxia-inducible factor 2alpha plays an important role in vascular remodeling. Proc. Natl. Acad. Sci. USA 2000, 97, 8386–8391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Compernolle, V.; Brusselmans, K.; Acker, T.; Hoet, P.; Tjwa, M.; Beck, H.; Plaisance, S.; Dor, Y.; Keshet, E.; Lupu, F.; et al. Loss of HIF-2alpha and inhibition of VEGF impair fetal lung maturation, whereas treatment with VEGF prevents fatal respiratory distress in premature mice. Nat. Med. 2002, 8, 702–710. [Google Scholar] [CrossRef] [PubMed]
  36. Pan, H.; Chen, Q.; Qi, S.; Li, T.; Liu, B.; Liu, S.; Ma, X.; Wang, B. Mutations in EPAS1 in congenital heart disease in Tibetans. BioSci. Rep. 2018, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. An, J.Y.; Cristino, A.S.; Zhao, Q.; Edson, J.; Williams, S.M.; Ravine, D.; Wray, J.; Marshall, V.M.; Hunt, A.; Whitehouse, A.J.; et al. Towards a molecular characterization of autism spectrum disorders: An exome sequencing and systems approach. Transl. Psychiatry 2014, 4, e394. [Google Scholar] [CrossRef] [Green Version]
  38. Lorenzo, F.R.; Yang, C.; Ng Tang Fui, M.; Vankayalapati, H.; Zhuang, Z.; Huynh, T.; Grossmann, M.; Pacak, K.; Prchal, J.T. A novel EPAS1/HIF2A germline mutation in a congenital polycythemia with paraganglioma. J. Mol. Med. 2013, 91, 507–512. [Google Scholar] [CrossRef] [Green Version]
  39. Welander, J.; Andreasson, A.; Brauckhoff, M.; Backdahl, M.; Larsson, C.; Gimm, O.; Soderkvist, P. Frequent EPAS1/HIF2alpha exons 9 and 12 mutations in non-familial pheochromocytoma. Endocr. Relat. Cancer 2014, 21, 495–504. [Google Scholar] [CrossRef] [Green Version]
  40. Yang, C.; Hong, C.S.; Prchal, J.T.; Balint, M.T.; Pacak, K.; Zhuang, Z. Somatic mosaicism of EPAS1 mutations in the syndrome of paraganglioma and somatostatinoma associated with polycythemia. Hum. Genome. Var. 2015, 2, 15053. [Google Scholar] [CrossRef] [Green Version]
  41. Comino-Mendez, I.; de Cubas, A.A.; Bernal, C.; Alvarez-Escola, C.; Sanchez-Malo, C.; Ramirez-Tortosa, C.L.; Pedrinaci, S.; Rapizzi, E.; Ercolino, T.; Bernini, G.; et al. Tumoral EPAS1 (HIF2A) mutations explain sporadic pheochromocytoma and paraganglioma in the absence of erythrocytosis. Hum. Mol. Genet. 2013, 22, 2169–2176. [Google Scholar] [CrossRef] [PubMed]
  42. Zhuang, Z.; Yang, C.; Lorenzo, F.; Merino, M.; Fojo, T.; Kebebew, E.; Popovic, V.; Stratakis, C.A.; Prchal, J.T.; Pacak, K. Somatic HIF2A gain-of-function mutations in paraganglioma with polycythemia. N. Engl. J. Med. 2012, 367, 922–930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Toledo, R.A.; Qin, Y.; Srikantan, S.; Morales, N.P.; Li, Q.; Deng, Y.; Kim, S.W.; Pereira, M.A.; Toledo, S.P.; Su, X.; et al. In vivo and in vitro oncogenic effects of HIF2A mutations in pheochromocytomas and paragangliomas. Endocr. Relat. Cancer 2013, 20, 349–359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Buffet, A.; Smati, S.; Mansuy, L.; Menara, M.; Lebras, M.; Heymann, M.F.; Simian, C.; Favier, J.; Murat, A.; Cariou, B.; et al. Mosaicism in HIF2A-related polycythemia-paraganglioma syndrome. J. Clin. Endocrinol. Metab. 2014, 99, E369–E373. [Google Scholar] [CrossRef] [Green Version]
  45. Xiang, K.; Ouzhuluobu; Peng, Y.; Yang, Z.; Zhang, X.; Cui, C.; Zhang, H.; Li, M.; Zhang, Y.; Bianba; et al. Identification of a Tibetan-specific mutation in the hypoxic gene EGLN1 and its contribution to high-altitude adaptation. Mol. Biol. Evol. 2013, 30, 1889–1898. [Google Scholar] [CrossRef] [Green Version]
  46. Perrotta, S.; Stiehl, D.P.; Punzo, F.; Scianguetta, S.; Borriello, A.; Bencivenga, D.; Casale, M.; Nobili, B.; Fasoli, S.; Balduzzi, A.; et al. Congenital erythrocytosis associated with gain-of-function HIF2A gene mutations and erythropoietin levels in the normal range. Haematologica 2013, 98, 1624–1632. [Google Scholar] [CrossRef] [PubMed]
  47. Furlow, P.W.; Percy, M.J.; Sutherland, S.; Bierl, C.; McMullin, M.F.; Master, S.R.; Lappin, T.R.; Lee, F.S. Erythrocytosis-associated HIF-2alpha mutations demonstrate a critical role for residues C-terminal to the hydroxylacceptor proline. J. Biol. Chem. 2009, 284, 9050–9058. [Google Scholar] [CrossRef] [Green Version]
  48. Percy, M.J.; Chung, Y.J.; Harrison, C.; Mercieca, J.; Hoffbrand, A.V.; Dinardo, C.L.; Santos, P.C.; Fonseca, G.H.; Gualandro, S.F.; Pereira, A.C.; et al. Two new mutations in the HIF2A gene associated with erythrocytosis. Am. J. Hematol. 2012, 87, 439–442. [Google Scholar] [CrossRef] [Green Version]
  49. Martini, M.; Teofili, L.; Cenci, T.; Giona, F.; Torti, L.; Rea, M.; Foa, R.; Leone, G.; Larocca, L.M. A novel heterozygous HIF2AM535I mutation reinforces the role of oxygen sensing pathway disturbances in the pathogenesis of familial erythrocytosis. Haematologica 2008, 93, 1068–1071. [Google Scholar] [CrossRef] [Green Version]
  50. Percy, M.J.; Beer, P.A.; Campbell, G.; Dekker, A.W.; Green, A.R.; Oscier, D.; Rainey, M.G.; van Wijk, R.; Wood, M.; Lappin, T.R.; et al. Novel exon 12 mutations in the HIF2A gene associated with erythrocytosis. Blood 2008, 111, 5400–5402. [Google Scholar] [CrossRef]
  51. van Wijk, R.; Sutherland, S.; Van Wesel, A.C.; Huizinga, E.G.; Percy, M.J.; Bierings, M.; Lee, F.S. Erythrocytosis associated with a novel missense mutation in the HIF2A gene. Haematologica 2010, 95, 829–832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Gale, D.P.; Harten, S.K.; Reid, C.D.; Tuddenham, E.G.; Maxwell, P.H. Autosomal dominant erythrocytosis and pulmonary arterial hypertension associated with an activating HIF2 alpha mutation. Blood 2008, 112, 919–921. [Google Scholar] [CrossRef] [PubMed]
  53. Yamashita, T.; Ohneda, O.; Nagano, M.; Iemitsu, M.; Makino, Y.; Tanaka, H.; Miyauchi, T.; Goto, K.; Ohneda, K.; Fujii-Kuriyama, Y.; et al. Abnormal heart development and lung remodeling in mice lacking the hypoxia-inducible factor-related basic helix-loop-helix PAS protein NEPAS. Mol. Cell. Biol. 2008, 28, 1285–1297. [Google Scholar] [CrossRef] [Green Version]
  54. Takeda, K.; Ho, V.C.; Takeda, H.; Duan, L.J.; Nagy, A.; Fong, G.H. Placental but not heart defects are associated with elevated hypoxia-inducible factor alpha levels in mice lacking prolyl hydroxylase domain protein 2. Mol. Cell. Biol. 2006, 26, 8336–8346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Zhu, Z.; Gao, X.; He, Y.; Zhao, H.; Yu, Q.; Jiang, D.; Zhang, P.; Ma, X.; Huang, H.; Dong, D.; et al. An insertion/deletion polymorphism within RERT-lncRNA modulates hepatocellular carcinoma risk. Cancer Res. 2012, 72, 6163–6172. [Google Scholar] [CrossRef] [Green Version]
  56. Che, J.; Jiang, D.; Zheng, Y.; Zhu, B.; Zhang, P.; Lu, D.; Zhang, J.; Xiao, J.; Wang, J.; Gao, Y.; et al. Polymorphism in PHD1 gene and risk of non-small cell lung cancer in a Chinese population. Tumour. Biol. 2014, 35, 8921–8925. [Google Scholar] [CrossRef]
  57. Zhu, J.; Luo, J.Z.; Li, C.B. Correlations of an Insertion/Deletion Polymorphism (rs10680577) in the RERT-lncRNA with the Susceptibility, Clinicopathological Features, and Prognosis of Lung Cancer. Biochem. Genet. 2019, 57, 147–158. [Google Scholar] [CrossRef]
  58. Wang, L.; Yamaguchi, S.; Burstein, M.D.; Terashima, K.; Chang, K.; Ng, H.K.; Nakamura, H.; He, Z.; Doddapaneni, H.; Lewis, L.; et al. Novel somatic and germline mutations in intracranial germ cell tumours. Nature 2014, 511, 241–245. [Google Scholar] [CrossRef] [Green Version]
  59. Li, C.; Feng, L.; Niu, L.; Teng Li, T.; Zhang, B.; Wan, H.; Zhu, Z.; Liu, H.; Wang, K.; Fu, H.; et al. An insertion/deletion polymorphism within the promoter of EGLN2 is associated with susceptibility to colorectal cancer. Int. J. Biol. Markers 2017, 32, e274–e277. [Google Scholar] [CrossRef] [Green Version]
  60. Yang, C.; Zhuang, Z.; Fliedner, S.M.; Shankavaram, U.; Sun, M.G.; Bullova, P.; Zhu, R.; Elkahloun, A.G.; Kourlas, P.J.; Merino, M.; et al. Germ-line PHD1 and PHD2 mutations detected in patients with pheochromocytoma/paraganglioma-polycythemia. J. Mol. Med. 2015, 93, 93–104. [Google Scholar] [CrossRef]
  61. Ladroue, C.; Hoogewijs, D.; Gad, S.; Carcenac, R.; Storti, F.; Barrois, M.; Gimenez-Roqueplo, A.P.; Leporrier, M.; Casadevall, N.; Hermine, O.; et al. Distinct deregulation of the hypoxia inducible factor by PHD2 mutants identified in germline DNA of patients with polycythemia. Haematologica 2012, 97, 9–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Albiero, E.; Ruggeri, M.; Fortuna, S.; Bernardi, M.; Finotto, S.; Madeo, D.; Rodeghiero, F. Analysis of the oxygen sensing pathway genes in familial chronic myeloproliferative neoplasms and identification of a novel EGLN1 germ-line mutation. Br. J. Haematol. 2011, 153, 405–408. [Google Scholar] [CrossRef] [PubMed]
  63. Al-Sheikh, M.; Moradkhani, K.; Lopez, M.; Wajcman, H.; Prehu, C. Disturbance in the HIF-1alpha pathway associated with erythrocytosis: Further evidences brought by frameshift and nonsense mutations in the prolyl hydroxylase domain protein 2 (PHD2) gene. Blood Cells Mol. Dis. 2008, 40, 160–165. [Google Scholar] [CrossRef]
  64. Bento, C.; Percy, M.J.; Gardie, B.; Maia, T.M.; van Wijk, R.; Perrotta, S.; Della Ragione, F.; Almeida, H.; Rossi, C.; Girodon, F.; et al. Genetic basis of congenital erythrocytosis: Mutation update and online databases. Hum. Mutat. 2014, 35, 15–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Jang, J.H.; Seo, J.Y.; Jang, J.; Jung, C.W.; Lee, K.O.; Kim, S.H.; Kim, H.J. Hereditary gene mutations in Korean patients with isolated erythrocytosis. Ann. Hematol. 2014, 93, 931–935. [Google Scholar] [CrossRef] [PubMed]
  66. Percy, M.J.; Zhao, Q.; Flores, A.; Harrison, C.; Lappin, T.R.; Maxwell, P.H.; McMullin, M.F.; Lee, F.S. A family with erythrocytosis establishes a role for prolyl hydroxylase domain protein 2 in oxygen homeostasis. Proc. Natl. Acad. Sci. USA 2006, 103, 654–659. [Google Scholar] [CrossRef] [Green Version]
  67. Bento, C.; Almeida, H.; Maia, T.M.; Relvas, L.; Oliveira, A.C.; Rossi, C.; Girodon, F.; Fernandez-Lago, C.; Aguado-Diaz, A.; Fraga, C.; et al. Molecular study of congenital erythrocytosis in 70 unrelated patients revealed a potential causal mutation in less than half of the cases (Where is/are the missing gene(s)?). Eur. J. Haematol. 2013, 91, 361–368. [Google Scholar] [CrossRef] [Green Version]
  68. Wilson, R.; Syed, N.; Shah, P. Erythrocytosis due to PHD2 Mutations: A Review of Clinical Presentation, Diagnosis, and Genetics. Case Rep. Hematol. 2016, 2016, 6373706. [Google Scholar] [CrossRef] [Green Version]
  69. Percy, M.J.; Furlow, P.W.; Beer, P.A.; Lappin, T.R.; McMullin, M.F.; Lee, F.S. A novel erythrocytosis-associated PHD2 mutation suggests the location of a HIF binding groove. Blood 2007, 110, 2193–2196. [Google Scholar] [CrossRef] [Green Version]
  70. Ladroue, C.; Carcenac, R.; Leporrier, M.; Gad, S.; Le Hello, C.; Galateau-Salle, F.; Feunteun, J.; Pouyssegur, J.; Richard, S.; Gardie, B. PHD2 mutation and congenital erythrocytosis with paraganglioma. N. Engl. J. Med. 2008, 359, 2685–2692. [Google Scholar] [CrossRef]
  71. Albiero, E.; Ruggeri, M.; Fortuna, S.; Finotto, S.; Bernardi, M.; Madeo, D.; Rodeghiero, F. Isolated erythrocytosis: Study of 67 patients and identification of three novel germ-line mutations in the prolyl hydroxylase domain protein 2 (PHD2) gene. Haematologica 2012, 97, 123–127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Talbot, N.P.; Smith, T.G.; Balanos, G.M.; Dorrington, K.L.; Maxwell, P.H.; Robbins, P.A. Cardiopulmonary phenotype associated with human PHD2 mutation. Physiol. Rep. 2017, 5. [Google Scholar] [CrossRef] [PubMed]
  73. Aragones, J.; Schneider, M.; Van Geyte, K.; Fraisl, P.; Dresselaers, T.; Mazzone, M.; Dirkx, R.; Zacchigna, S.; Lemieux, H.; Jeoung, N.H.; et al. Deficiency or inhibition of oxygen sensor Phd1 induces hypoxia tolerance by reprogramming basal metabolism. Nat. Genet. 2008, 40, 170–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Holster, T.; Pakkanen, O.; Soininen, R.; Sormunen, R.; Nokelainen, M.; Kivirikko, K.I.; Myllyharju, J. Loss of assembly of the main basement membrane collagen, type IV, but not fibril-forming collagens and embryonic death in collagen prolyl 4-hydroxylase I null mice. J. Biol. Chem. 2007, 282, 2512–2519. [Google Scholar] [CrossRef] [Green Version]
  75. Zou, Y.; Donkervoort, S.; Salo, A.M.; Foley, A.R.; Barnes, A.M.; Hu, Y.; Makareeva, E.; Leach, M.E.; Mohassel, P.; Dastgir, J.; et al. P4HA1 mutations cause a unique congenital disorder of connective tissue involving tendon, bone, muscle and the eye. Hum. Mol. Genet. 2017, 26, 2207–2217. [Google Scholar] [CrossRef] [Green Version]
  76. Aro, E.; Salo, A.M.; Khatri, R.; Finnila, M.; Miinalainen, I.; Sormunen, R.; Pakkanen, O.; Holster, T.; Soininen, R.; Prein, C.; et al. Severe Extracellular Matrix Abnormalities and Chondrodysplasia in Mice Lacking Collagen Prolyl 4-Hydroxylase Isoenzyme II in Combination with a Reduced Amount of Isoenzyme I. J. Biol. Chem. 2015, 290, 16964–16978. [Google Scholar] [CrossRef] [Green Version]
  77. Guo, H.; Tong, P.; Liu, Y.; Xia, L.; Wang, T.; Tian, Q.; Li, Y.; Hu, Y.; Zheng, Y.; Jin, X.; et al. Mutations of P4HA2 encoding prolyl 4-hydroxylase 2 are associated with nonsyndromic high myopia. Genet. Med. 2015, 17, 300–306. [Google Scholar] [CrossRef] [Green Version]
  78. Ferdinandusse, S.; Zomer, A.W.; Komen, J.C.; van den Brink, C.E.; Thanos, M.; Hamers, F.P.; Wanders, R.J.; van der Saag, P.T.; Poll-The, B.T.; Brites, P. Ataxia with loss of Purkinje cells in a mouse model for Refsum disease. Proc. Natl. Acad. Sci. USA 2008, 105, 17712–17717. [Google Scholar] [CrossRef] [Green Version]
  79. Jansen, G.A.; Hogenhout, E.M.; Ferdinandusse, S.; Waterham, H.R.; Ofman, R.; Jakobs, C.; Skjeldal, O.H.; Wanders, R.J. Human phytanoyl-CoA hydroxylase: Resolution of the gene structure and the molecular basis of Refsum’s disease. Hum. Mol. Genet. 2000, 9, 1195–1200. [Google Scholar] [CrossRef] [Green Version]
  80. Chahal, A.; Khan, M.; Pai, S.G.; Barbosa, E.; Singh, I. Restoration of phytanic acid oxidation in Refsum disease fibroblasts from patients with mutations in the phytanoyl-CoA hydroxylase gene. FEBS Lett. 1998, 429, 119–122. [Google Scholar] [CrossRef] [Green Version]
  81. Jansen, G.A.; Waterham, H.R.; Wanders, R.J. Molecular basis of Refsum disease: Sequence variations in phytanoyl-CoA hydroxylase (PHYH) and the PTS2 receptor (PEX7). Hum. Mutat. 2004, 23, 209–218. [Google Scholar] [CrossRef] [PubMed]
  82. Jansen, G.A.; Ferdinandusse, S.; Skjeldal, O.H.; Stokke, O.; de Groot, C.J.; Jakobs, C.; Wanders, R.J. Molecular basis of Refsum disease: Identification of new mutations in the phytanoyl-CoA hydroxylase cDNA. J. Inherit. Metab. Dis. 1998, 21, 288–291. [Google Scholar] [CrossRef] [PubMed]
  83. Zhao, L.; Wang, F.; Wang, H.; Li, Y.; Alexander, S.; Wang, K.; Willoughby, C.E.; Zaneveld, J.E.; Jiang, L.; Soens, Z.T.; et al. Next-generation sequencing-based molecular diagnosis of 82 retinitis pigmentosa probands from Northern Ireland. Hum. Genet. 2015, 134, 217–230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Mihalik, S.J.; Morrell, J.C.; Kim, D.; Sacksteder, K.A.; Watkins, P.A.; Gould, S.J. Identification of PAHX, a Refsum disease gene. Nat. Genet. 1997, 17, 185–189. [Google Scholar] [CrossRef] [PubMed]
  85. Kohlschutter, A.; Santer, R.; Lukacs, Z.; Altenburg, C.; Kemper, M.J.; Ruther, K. A child with night blindness: Preventing serious symptoms of Refsum disease. J. Child Neurol. 2012, 27, 654–656. [Google Scholar] [CrossRef] [PubMed]
  86. Aylward, A.; Cai, Y.; Lee, A.; Blue, E.; Rabinowitz, D.; Haddad, J., Jr.; University of Washington Center for Mendelian Genomics. Using Whole Exome Sequencing to Identify Candidate Genes With Rare Variants In Nonsyndromic Cleft Lip and Palate. Genet. Epidemiol. 2016, 40, 432–441. [Google Scholar] [CrossRef]
  87. Zhang, N.; Fu, Z.; Linke, S.; Chicher, J.; Gorman, J.J.; Visk, D.; Haddad, G.G.; Poellinger, L.; Peet, D.J.; Powell, F.; et al. The asparaginyl hydroxylase factor inhibiting HIF-1alpha is an essential regulator of metabolism. Cell Metab. 2010, 11, 364–378. [Google Scholar] [CrossRef] [Green Version]
  88. Webb, E.L.; Rudd, M.F.; Sellick, G.S.; El Galta, R.; Bethke, L.; Wood, W.; Fletcher, O.; Penegar, S.; Withey, L.; Qureshi, M.; et al. Search for low penetrance alleles for colorectal cancer through a scan of 1467 non-synonymous SNPs in 2575 cases and 2707 controls with validation by kin-cohort analysis of 14 704 first-degree relatives. Hum. Mol. Genet. 2006, 15, 3263–3271. [Google Scholar] [CrossRef] [Green Version]
  89. Yamaguchi, S.; Hong, K.; Liu, R.; Shen, L.; Inoue, A.; Diep, D.; Zhang, K.; Zhang, Y. Tet1 controls meiosis by regulating meiotic gene expression. Nature 2012, 492, 443–447. [Google Scholar] [CrossRef] [Green Version]
  90. Kang, J.; Lienhard, M.; Pastor, W.A.; Chawla, A.; Novotny, M.; Tsagaratou, A.; Lasken, R.S.; Thompson, E.C.; Surani, M.A.; Koralov, S.B.; et al. Simultaneous deletion of the methylcytosine oxidases Tet1 and Tet3 increases transcriptome variability in early embryogenesis. Proc. Natl. Acad. Sci. USA 2015, 112, E4236–E4245. [Google Scholar] [CrossRef] [Green Version]
  91. Khoueiry, R.; Sohni, A.; Thienpont, B.; Luo, X.; Velde, J.V.; Bartoccetti, M.; Boeckx, B.; Zwijsen, A.; Rao, A.; Lambrechts, D.; et al. Lineage-specific functions of TET1 in the postimplantation mouse embryo. Nat. Genet. 2017, 49, 1061–1072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Dawlaty, M.M.; Ganz, K.; Powell, B.E.; Hu, Y.C.; Markoulaki, S.; Cheng, A.W.; Gao, Q.; Kim, J.; Choi, S.W.; Page, D.C.; et al. Tet1 is dispensable for maintaining pluripotency and its loss is compatible with embryonic and postnatal development. Cell Stem Cell 2011, 9, 166–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Liu, J.; Yue, Y.; Han, D.; Wang, X.; Fu, Y.; Zhang, L.; Jia, G.; Yu, M.; Lu, Z.; Deng, X.; et al. A METTL3-METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation. Nat. Chem. Biol. 2014, 10, 93–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zhang, R.R.; Cui, Q.Y.; Murai, K.; Lim, Y.C.; Smith, Z.D.; Jin, S.; Ye, P.; Rosa, L.; Lee, Y.K.; Wu, H.P.; et al. Tet1 regulates adult hippocampal neurogenesis and cognition. Cell Stem Cell 2013, 13, 237–245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Li, Z.; Cai, X.; Cai, C.L.; Wang, J.; Zhang, W.; Petersen, B.E.; Yang, F.C.; Xu, M. Deletion of Tet2 in mice leads to dysregulated hematopoietic stem cells and subsequent development of myeloid malignancies. Blood 2011, 118, 4509–4518. [Google Scholar] [CrossRef] [Green Version]
  96. Pan, F.; Wingo, T.S.; Zhao, Z.; Gao, R.; Makishima, H.; Qu, G.; Lin, L.; Yu, M.; Ortega, J.R.; Wang, J.; et al. Tet2 loss leads to hypermutagenicity in haematopoietic stem/progenitor cells. Nat. Commun. 2017, 8, 15102. [Google Scholar] [CrossRef] [Green Version]
  97. Zhao, Z.; Chen, L.; Dawlaty, M.M.; Pan, F.; Weeks, O.; Zhou, Y.; Cao, Z.; Shi, H.; Wang, J.; Lin, L.; et al. Combined Loss of Tet1 and Tet2 Promotes B Cell, but Not Myeloid Malignancies, in Mice. Cell Rep. 2015, 13, 1692–1704. [Google Scholar] [CrossRef] [Green Version]
  98. Ismael, O.; Shimada, A.; Hama, A.; Elshazley, M.; Muramatsu, H.; Goto, A.; Sakaguchi, H.; Tanaka, M.; Takahashi, Y.; Yinyan, X.; et al. De novo childhood myelodysplastic/myeloproliferative disease with unique molecular characteristics. Br. J. Haematol. 2012, 158, 129–137. [Google Scholar] [CrossRef]
  99. Nickerson, M.L.; Im, K.M.; Misner, K.J.; Tan, W.; Lou, H.; Gold, B.; Wells, D.W.; Bravo, H.C.; Fredrikson, K.M.; Harkins, T.T.; et al. Somatic alterations contributing to metastasis of a castration-resistant prostate cancer. Hum. Mutat. 2013, 34, 1231–1241. [Google Scholar] [CrossRef] [Green Version]
  100. Schaub, F.X.; Looser, R.; Li, S.; Hao-Shen, H.; Lehmann, T.; Tichelli, A.; Skoda, R.C. Clonal analysis of TET2 and JAK2 mutations suggests that TET2 can be a late event in the progression of myeloproliferative neoplasms. Blood 2010, 115, 2003–2007. [Google Scholar] [CrossRef] [Green Version]
  101. Gu, T.P.; Guo, F.; Yang, H.; Wu, H.P.; Xu, G.F.; Liu, W.; Xie, Z.G.; Shi, L.; He, X.; Jin, S.G.; et al. The role of tet3 DNA dioxygenase in epigenetic reprogramming by oocytes. Nature 2011, 477, 606–610. [Google Scholar] [CrossRef]
  102. Beck, D.B.; Petracovici, A.; He, C.; Moore, H.W.; Louie, R.J.; Ansar, M.; Douzgou, S.; Sithambaram, S.; Cottrell, T.; Santos-Cortez, R.L.P.; et al. Delineation of a Human Mendelian Disorder of the DNA Demethylation Machinery: TET3 Deficiency. Am. J. Hum. Genet. 2020, 106, 234–245. [Google Scholar] [CrossRef]
  103. Boissel, S.; Reish, O.; Proulx, K.; Kawagoe-Takaki, H.; Sedgwick, B.; Yeo, G.S.; 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] [Green Version]
  104. Caglayan, A.O.; Tuysuz, B.; Coskun, S.; Quon, J.; Harmanci, A.S.; Baranoski, J.F.; Baran, B.; Erson-Omay, E.Z.; Henegariu, O.; Mane, S.M.; et al. A patient with a novel homozygous missense mutation in FTO and concomitant nonsense mutation in CETP. J. Hum. Genet. 2016, 61, 395–403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. 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. 2016, 53, 200–207. [Google Scholar] [CrossRef] [PubMed]
  106. Rohena, L.; Lawson, M.; Guzman, E.; Ganapathi, M.; Cho, M.T.; Haverfield, E.; Anyane-Yeboa, K. FTO variant associated with malformation syndrome. Am. J. Med. Genet. A 2016, 170A, 1023–1028. [Google Scholar] [CrossRef] [PubMed]
  107. Song, Y.; You, N.C.; Hsu, Y.H.; Howard, B.V.; Langer, R.D.; Manson, J.E.; Nathan, L.; Niu, T.; Tinker, L.F.; Liu, S. FTO polymorphisms are associated with obesity but not diabetes risk in postmenopausal women. Obesity 2008, 16, 2472–2480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Yang, J.; Loos, R.J.; Powell, J.E.; Medland, S.E.; Speliotes, E.K.; Chasman, D.I.; Rose, L.M.; Thorleifsson, G.; Steinthorsdottir, V.; Magi, R.; et al. FTO genotype is associated with phenotypic variability of body mass index. Nature 2012, 490, 267–272. [Google Scholar] [CrossRef] [Green Version]
  109. Frayling, T.M.; Timpson, N.J.; Weedon, M.N.; Zeggini, E.; Freathy, R.M.; Lindgren, C.M.; Perry, J.R.; 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] [Green Version]
  110. Hotta, K.; Kitamoto, T.; Kitamoto, A.; Mizusawa, S.; Matsuo, T.; Nakata, Y.; Kamohara, S.; Miyatake, N.; Kotani, K.; Komatsu, R.; et al. Association of variations in the FTO, SCG3 and MTMR9 genes with metabolic syndrome in a Japanese population. J. Hum. Genet. 2011, 56, 647–651. [Google Scholar] [CrossRef] [Green Version]
  111. Yoo, H.; Son, D.; Lee, Y.J.; Hong, K. Mouse JMJD4 is dispensable for embryogenesis. Mol. Reprod. Dev. 2016, 83, 588–593. [Google Scholar] [CrossRef] [PubMed]
  112. Li, M.O.; Sarkisian, M.R.; Mehal, W.Z.; Rakic, P.; Flavell, R.A. Phosphatidylserine receptor is required for clearance of apoptotic cells. Science 2003, 302, 1560–1563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Kunisaki, Y.; Masuko, S.; Noda, M.; Inayoshi, A.; Sanui, T.; Harada, M.; Sasazuki, T.; Fukui, Y. Defective fetal liver erythropoiesis and T lymphopoiesis in mice lacking the phosphatidylserine receptor. Blood 2004, 103, 3362–3364. [Google Scholar] [CrossRef] [PubMed]
  114. Bose, J.; Gruber, A.D.; Helming, L.; Schiebe, S.; Wegener, I.; Hafner, M.; Beales, M.; Kontgen, F.; Lengeling, A. The phosphatidylserine receptor has essential functions during embryogenesis but not in apoptotic cell removal. J. Biol. 2004, 3, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Schneider, J.E.; Bose, J.; Bamforth, S.D.; Gruber, A.D.; Broadbent, C.; Clarke, K.; Neubauer, S.; Lengeling, A.; Bhattacharya, S. Identification of cardiac malformations in mice lacking Ptdsr using a novel high-throughput magnetic resonance imaging technique. BMC Dev. Biol. 2004, 4, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Kawakami, E.; Tokunaga, A.; Ozawa, M.; Sakamoto, R.; Yoshida, N. The histone demethylase Fbxl11/Kdm2a plays an essential role in embryonic development by repressing cell-cycle regulators. Mech. Mech. Dev.. 2015, 135, 31–42. [Google Scholar] [CrossRef]
  117. Fukuda, T.; Tokunaga, A.; Sakamoto, R.; Yoshida, N. Fbxl10/Kdm2b deficiency accelerates neural progenitor cell death and leads to exencephaly. Mol. Cell. Neurosci. 2011, 46, 614–624. [Google Scholar] [CrossRef]
  118. Boulard, M.; Edwards, J.R.; Bestor, T.H. FBXL10 protects Polycomb-bound genes from hypermethylation. Nat. Genet. 2015, 47, 479–485. [Google Scholar] [CrossRef]
  119. Boulard, M.; Edwards, J.R.; Bestor, T.H. Abnormal X chromosome inactivation and sex-specific gene dysregulation after ablation of FBXL10. Epigenet. Chromatin 2016, 9, 22. [Google Scholar] [CrossRef] [Green Version]
  120. Andricovich, J.; Kai, Y.; Peng, W.; Foudi, A.; Tzatsos, A. Histone demethylase KDM2B regulates lineage commitment in normal and malignant hematopoiesis. J. Clin. Investig. 2016, 126, 905–920. [Google Scholar] [CrossRef]
  121. Tateishi, K.; Okada, Y.; Kallin, E.M.; Zhang, Y. Role of Jhdm2a in regulating metabolic gene expression and obesity resistance. Nature 2009, 458, 757–761. [Google Scholar] [CrossRef] [Green Version]
  122. Inagaki, T.; Tachibana, M.; Magoori, K.; Kudo, H.; Tanaka, T.; Okamura, M.; Naito, M.; Kodama, T.; Shinkai, Y.; Sakai, J. Obesity and metabolic syndrome in histone demethylase JHDM2a-deficient mice. Genes Cells 2009, 14, 991–1001. [Google Scholar] [CrossRef]
  123. Liu, Z.; Zhou, S.; Liao, L.; Chen, X.; Meistrich, M.; Xu, J. Jmjd1a demethylase-regulated histone modification is essential for cAMP-response element modulator-regulated gene expression and spermatogenesis. J. Biol. Chem. 2010, 285, 2758–2770. [Google Scholar] [CrossRef] [Green Version]
  124. Qin, L.; Xu, Y.; Yu, X.; Toneff, M.J.; Li, D.; Liao, L.; Martinez, J.D.; Li, Y.; Xu, J. The histone demethylase Kdm3a is required for normal epithelial proliferation, ductal elongation and tumor growth in the mouse mammary gland. Oncotarget 2017, 8, 84761–84775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Hojati, Z.; Soleimanpour, E.; Javadirad, S.M.; Nasr-Esfahani, M.H. Identification of Two Novel Mutations in KDM3A Regulatory Gene in Iranian Infertile Males. Iran. Biomed. J. 2019, 23, 220–227. [Google Scholar] [PubMed]
  126. Liu, Z.; Chen, X.; Zhou, S.; Liao, L.; Jiang, R.; Xu, J. The histone H3K9 demethylase Kdm3b is required for somatic growth and female reproductive function. J. Biol.Int. J. Biol. Sci. 2015, 11, 494–507. [Google Scholar] [CrossRef] [PubMed]
  127. Li, S.; Ali, S.; Duan, X.; Liu, S.; Du, J.; Liu, C.; Dai, H.; Zhou, M.; Zhou, L.; Yang, L.; et al. JMJD1B Demethylates H4R3me2s and H3K9me2 to Facilitate Gene Expression for Development of Hematopoietic Stem and Progenitor Cells. Cell Rep. 2018, 23, 389–403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Guipponi, M.; Santoni, F.A.; Setola, V.; Gehrig, C.; Rotharmel, M.; Cuenca, M.; Guillin, O.; Dikeos, D.; Georgantopoulos, G.; Papadimitriou, G.; et al. Exome sequencing in 53 sporadic cases of schizophrenia identifies 18 putative candidate genes. PLoS ONE 2014, 9, e112745. [Google Scholar] [CrossRef]
  129. Deciphering Developmental Disorders, S. Prevalence and architecture of de novo mutations in developmental disorders. Nature 2017, 542, 433–438. [Google Scholar] [CrossRef]
  130. Mahamdallie, S.; Yost, S.; Poyastro-Pearson, E.; Holt, E.; Zachariou, A.; Seal, S.; Elliott, A.; Clarke, M.; Warren-Perry, M.; Hanks, S.; et al. Identification of new Wilms tumour predisposition genes: An exome sequencing study. Lancet Child Adolesc. Health 2019, 3, 322–331. [Google Scholar] [CrossRef] [Green Version]
  131. Diets, I.J.; van der Donk, R.; Baltrunaite, K.; Waanders, E.; Reijnders, M.R.F.; Dingemans, A.J.M.; Pfundt, R.; Vulto-van Silfhout, A.T.; Wiel, L.; Gilissen, C.; et al. De Novo and Inherited Pathogenic Variants in KDM3B Cause Intellectual Disability, Short Stature, and Facial Dysmorphism. Am. J. Hum. Genet. 2019, 104, 758–766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Guo, T.; Chung, J.H.; Wang, T.; McDonald-McGinn, D.M.; Kates, W.R.; Hawula, W.; Coleman, K.; Zackai, E.; Emanuel, B.S.; Morrow, B.E. Histone Modifier Genes Alter Conotruncal Heart Phenotypes in 22q11.2 Deletion Syndrome. Am. J. Hum. Genet. 2015, 97, 869–877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Saez, M.A.; Fernandez-Rodriguez, J.; Moutinho, C.; Sanchez-Mut, J.V.; Gomez, A.; Vidal, E.; Petazzi, P.; Szczesna, K.; Lopez-Serra, P.; Lucariello, M.; et al. Mutations in JMJD1C are involved in Rett syndrome and intellectual disability. Genet. Med. 2016, 18, 378–385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Neale, B.M.; Kou, Y.; Liu, L.; Ma’ayan, A.; Samocha, K.E.; Sabo, A.; Lin, C.F.; Stevens, C.; Wang, L.S.; Makarov, V.; et al. Patterns and rates of exonic de novo mutations in autism spectrum disorders. Nature 2012, 485, 242–245. [Google Scholar] [CrossRef] [PubMed]
  135. Zhang, Q.J.; Chen, H.Z.; Wang, L.; Liu, D.P.; Hill, J.A.; Liu, Z.P. The histone trimethyllysine demethylase JMJD2A promotes cardiac hypertrophy in response to hypertrophic stimuli in mice. J. Clin. Investig. 2011, 121, 2447–2456. [Google Scholar] [CrossRef] [Green Version]
  136. Kawazu, M.; Saso, K.; Tong, K.I.; McQuire, T.; Goto, K.; Son, D.O.; Wakeham, A.; Miyagishi, M.; Mak, T.W.; Okada, H. Histone demethylase JMJD2B functions as a co-factor of estrogen receptor in breast cancer proliferation and mammary gland development. PLoS ONE 2011, 6, e17830. [Google Scholar] [CrossRef]
  137. Cheng, Y.; Yuan, Q.; Vergnes, L.; Rong, X.; Youn, J.Y.; Li, J.; Yu, Y.; Liu, W.; Cai, H.; Lin, J.D.; et al. KDM4B protects against obesity and metabolic dysfunction. Proc. Natl. Acad. Sci. USA 2018, 115, E5566–E5575. [Google Scholar] [CrossRef] [Green Version]
  138. Pedersen, M.T.; Agger, K.; Laugesen, A.; Johansen, J.V.; Cloos, P.A.; Christensen, J.; Helin, K. The demethylase JMJD2C localizes to H3K4me3-positive transcription start sites and is dispensable for embryonic development. Mol. Cell. Biol. 2014, 34, 1031–1045. [Google Scholar] [CrossRef] [Green Version]
  139. Ozaki, Y.; Fujiwara, K.; Ikeda, M.; Ozaki, T.; Terui, T.; Soma, M.; Inazawa, J.; Nagase, H. The oncogenic role of GASC1 in chemically induced mouse skin cancer. Mamm. Genome 2015, 26, 591–597. [Google Scholar] [CrossRef]
  140. Canova, C.; Hashibe, M.; Simonato, L.; Nelis, M.; Metspalu, A.; Lagiou, P.; Trichopoulos, D.; Ahrens, W.; Pigeot, I.; Merletti, F.; et al. Genetic associations of 115 polymorphisms with cancers of the upper aerodigestive tract across 10 European countries: The ARCAGE project. Cancer Res. 2009, 69, 2956–2965. [Google Scholar] [CrossRef] [Green Version]
  141. Perry, J.R.; Day, F.; Elks, C.E.; Sulem, P.; Thompson, D.J.; Ferreira, T.; He, C.; Chasman, D.I.; Esko, T.; Thorleifsson, G.; et al. Parent-of-origin-specific allelic associations among 106 genomic loci for age at menarche. Nature 2014, 514, 92–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Iwamori, N.; Zhao, M.; Meistrich, M.L.; Matzuk, M.M. The testis-enriched histone demethylase, KDM4D, regulates methylation of histone H3 lysine 9 during spermatogenesis in the mouse but is dispensable for fertility. Biol. Reprod. 2011, 84, 1225–1234. [Google Scholar] [CrossRef] [PubMed]
  143. Klose, R.J.; Yan, Q.; Tothova, Z.; Yamane, K.; Erdjument-Bromage, H.; Tempst, P.; Gilliland, D.G.; Zhang, Y.; Kaelin, W.G., Jr. The retinoblastoma binding protein RBP2 is an H3K4 demethylase. Cell 2007, 128, 889–900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Lin, W.; Cao, J.; Liu, J.; Beshiri, M.L.; Fujiwara, Y.; Francis, J.; Cherniack, A.D.; Geisen, C.; Blair, L.P.; Zou, M.R.; et al. Loss of the retinoblastoma binding protein 2 (RBP2) histone demethylase suppresses tumorigenesis in mice lacking Rb1 or Men1. Proc. Natl. Acad. Sci. USA 2011, 108, 13379–13386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Najmabadi, H.; Hu, H.; Garshasbi, M.; Zemojtel, T.; Abedini, S.S.; Chen, W.; Hosseini, M.; Behjati, F.; Haas, S.; Jamali, P.; et al. Deep sequencing reveals 50 novel genes for recessive cognitive disorders. Nature 2011, 478, 57–63. [Google Scholar] [CrossRef]
  146. Zaidi, S.; Choi, M.; Wakimoto, H.; Ma, L.; Jiang, J.; Overton, J.D.; Romano-Adesman, A.; Bjornson, R.D.; Breitbart, R.E.; Brown, K.K.; et al. De novo mutations in histone-modifying genes in congenital heart disease. Nature 2013, 498, 220–223. [Google Scholar] [CrossRef] [Green Version]
  147. Catchpole, S.; Spencer-Dene, B.; Hall, D.; Santangelo, S.; Rosewell, I.; Guenatri, M.; Beatson, R.; Scibetta, A.G.; Burchell, J.M.; Taylor-Papadimitriou, J. PLU-1/JARID1B/KDM5B is required for embryonic survival and contributes to cell proliferation in the mammary gland and in ER+ breast cancer cells. Int. J. Oncol. 2011, 38, 1267–1277. [Google Scholar] [CrossRef] [Green Version]
  148. Zou, M.R.; Cao, J.; Liu, Z.; Huh, S.J.; Polyak, K.; Yan, Q. Histone demethylase jumonji AT-rich interactive domain 1B (JARID1B) controls mammary gland development by regulating key developmental and lineage specification genes. J. Biol. Chem. 2014, 289, 17620–17633. [Google Scholar] [CrossRef] [Green Version]
  149. Albert, M.; Schmitz, S.U.; Kooistra, S.M.; Malatesta, M.; Morales Torres, C.; Rekling, J.C.; Johansen, J.V.; Abarrategui, I.; Helin, K. The histone demethylase Jarid1b ensures faithful mouse development by protecting developmental genes from aberrant H3K4me3. PLoS Genet. 2013, 9, e1003461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Faundes, V.; Newman, W.G.; Bernardini, L.; Canham, N.; Clayton-Smith, J.; Dallapiccola, B.; Davies, S.J.; Demos, M.K.; Goldman, A.; Gill, H.; et al. Histone Lysine Methylases and Demethylases in the Landscape of Human Developmental Disorders. Am. J. Hum. Genet. 2018, 102, 175–187. [Google Scholar] [CrossRef] [PubMed]
  151. Cox, B.J.; Vollmer, M.; Tamplin, O.; Lu, M.; Biechele, S.; Gertsenstein, M.; van Campenhout, C.; Floss, T.; Kuhn, R.; Wurst, W.; et al. Phenotypic annotation of the mouse X chromosome. Genome Res. 2010, 20, 1154–1164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Iwase, S.; Brookes, E.; Agarwal, S.; Badeaux, A.I.; Ito, H.; Vallianatos, C.N.; Tomassy, G.S.; Kasza, T.; Lin, G.; Thompson, A.; et al. A Mouse Model of X-linked Intellectual Disability Associated with Impaired Removal of Histone Methylation. Cell Rep. 2016, 14, 1000–1009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Vallianatos, C.N.; Raines, B.; Porter, R.S.; Bonefas, K.M.; Wu, M.C.; Garay, P.M.; Collette, K.M.; Seo, Y.A.; Dou, Y.; Keegan, C.E.; et al. Mutually suppressive roles of KMT2A and KDM5C in behaviour, neuronal structure, and histone H3K4 methylation. Commun. Biol. 2020, 3, 278. [Google Scholar] [CrossRef] [PubMed]
  154. Ounap, K.; Puusepp-Benazzouz, H.; Peters, M.; Vaher, U.; Rein, R.; Proos, A.; Field, M.; Reimand, T. A novel c.2T > C mutation of the KDM5C/JARID1C gene in one large family with X-linked intellectual disability. Eur. J. Med. Genet. 2012, 55, 178–184. [Google Scholar] [CrossRef] [PubMed]
  155. Grozeva, D.; Carss, K.; Spasic-Boskovic, O.; Tejada, M.I.; Gecz, J.; Shaw, M.; Corbett, M.; Haan, E.; Thompson, E.; Friend, K.; et al. Targeted Next-Generation Sequencing Analysis of 1,000 Individuals with Intellectual Disability. Hum. Mutat. 2015, 36, 1197–1204. [Google Scholar] [CrossRef] [Green Version]
  156. Grafodatskaya, D.; Chung, B.H.; Butcher, D.T.; Turinsky, A.L.; Goodman, S.J.; Choufani, S.; Chen, Y.A.; Lou, Y.; Zhao, C.; Rajendram, R.; et al. Multilocus loss of DNA methylation in individuals with mutations in the histone H3 lysine 4 demethylase KDM5C. BMC Med. Genom. 2013, 6, 1. [Google Scholar] [CrossRef] [Green Version]
  157. Redin, C.; Gerard, B.; Lauer, J.; Herenger, Y.; Muller, J.; Quartier, A.; Masurel-Paulet, A.; Willems, M.; Lesca, G.; El-Chehadeh, S.; et al. Efficient strategy for the molecular diagnosis of intellectual disability using targeted high-throughput sequencing. J. Med. Genet. 2014, 51, 724–736. [Google Scholar] [CrossRef]
  158. Santos-Reboucas, C.B.; Fintelman-Rodrigues, N.; Jensen, L.R.; Kuss, A.W.; Ribeiro, M.G.; Campos, M., Jr.; Santos, J.M.; Pimentel, M.M. A novel nonsense mutation in KDM5C/JARID1C gene causing intellectual disability, short stature and speech delay. Neurosci. Lett. 2011, 498, 67–71. [Google Scholar] [CrossRef] [Green Version]
  159. Hu, H.; Haas, S.A.; Chelly, J.; Van Esch, H.; Raynaud, M.; de Brouwer, A.P.; Weinert, S.; Froyen, G.; Frints, S.G.; Laumonnier, F.; et al. X-exome sequencing of 405 unresolved families identifies seven novel intellectual disability genes. Mol. Psychiatry 2016, 21, 133–148. [Google Scholar] [CrossRef] [Green Version]
  160. Brookes, E.; Laurent, B.; Ounap, K.; Carroll, R.; Moeschler, J.B.; Field, M.; Schwartz, C.E.; Gecz, J.; Shi, Y. Mutations in the intellectual disability gene KDM5C reduce protein stability and demethylase activity. Hum. Mol. Genet. 2015, 24, 2861–2872. [Google Scholar] [CrossRef]
  161. Abidi, F.E.; Holloway, L.; Moore, C.A.; Weaver, D.D.; Simensen, R.J.; Stevenson, R.E.; Rogers, R.C.; Schwartz, C.E. Mutations in JARID1C are associated with X-linked mental retardation, short stature and hyperreflexia. J. Med. Genet. 2008, 45, 787–793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Tzschach, A.; Lenzner, S.; Moser, B.; Reinhardt, R.; Chelly, J.; Fryns, J.P.; Kleefstra, T.; Raynaud, M.; Turner, G.; Ropers, H.H.; et al. Novel JARID1C/SMCX mutations in patients with X-linked mental retardation. Hum. Mutat. 2006, 27, 389. [Google Scholar] [CrossRef] [PubMed]
  163. Jensen, L.R.; Amende, M.; Gurok, U.; Moser, B.; Gimmel, V.; Tzschach, A.; Janecke, A.R.; Tariverdian, G.; Chelly, J.; Fryns, J.P.; et al. Mutations in the JARID1C gene, which is involved in transcriptional regulation and chromatin remodeling, cause X-linked mental retardation. Am. J. Hum. Genet. 2005, 76, 227–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Santos, C.; Rodriguez-Revenga, L.; Madrigal, I.; Badenas, C.; Pineda, M.; Mila, M. A novel mutation in JARID1C gene associated with mental retardation. Eur. J. Hum. Genet. 2006, 14, 583–586. [Google Scholar] [CrossRef] [PubMed]
  165. Rujirabanjerd, S.; Nelson, J.; Tarpey, P.S.; Hackett, A.; Edkins, S.; Raymond, F.L.; Schwartz, C.E.; Turner, G.; Iwase, S.; Shi, Y.; et al. Identification and characterization of two novel JARID1C mutations: Suggestion of an emerging genotype-phenotype correlation. Eur. J. Hum. Genet. 2010, 18, 330–335. [Google Scholar] [CrossRef] [Green Version]
  166. Vissers, L.E.; de Ligt, J.; Gilissen, C.; Janssen, I.; Steehouwer, M.; de Vries, P.; van Lier, B.; Arts, P.; Wieskamp, N.; del Rosario, M.; et al. A de novo paradigm for mental retardation. Nat. Genet. 2010, 42, 1109–1112. [Google Scholar] [CrossRef]
  167. Adegbola, A.; Gao, H.; Sommer, S.; Browning, M. A novel mutation in JARID1C/SMCX in a patient with autism spectrum disorder (ASD). Am. J. Med. Genet. A 2008, 146A, 505–511. [Google Scholar] [CrossRef]
  168. Lu, Y.; Oura, S.; Matsumura, T.; Oji, A.; Sakurai, N.; Fujihara, Y.; Shimada, K.; Miyata, H.; Tobita, T.; Noda, T.; et al. CRISPR/Cas9-mediated genome editing reveals 30 testis-enriched genes dispensable for male fertility in micedagger. Biol. Reprod. 2019, 101, 501–511. [Google Scholar] [CrossRef]
  169. Shpargel, K.B.; Sengoku, T.; Yokoyama, S.; Magnuson, T. UTX and UTY demonstrate histone demethylase-independent function in mouse embryonic development. PLoS Genet. 2012, 8, e1002964. [Google Scholar] [CrossRef] [Green Version]
  170. Lee, S.; Lee, J.W.; Lee, S.K. UTX, a histone H3-lysine 27 demethylase, acts as a critical switch to activate the cardiac developmental program. Dev. Cell 2012, 22, 25–37. [Google Scholar] [CrossRef] [Green Version]
  171. Welstead, G.G.; Creyghton, M.P.; Bilodeau, S.; Cheng, A.W.; Markoulaki, S.; Young, R.A.; Jaenisch, R. X-linked H3K27me3 demethylase Utx is required for embryonic development in a sex-specific manner. Proc. Natl. Acad. Sci. USA 2012, 109, 13004–13009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Thieme, S.; Gyarfas, T.; Richter, C.; Ozhan, G.; Fu, J.; Alexopoulou, D.; Muders, M.H.; Michalk, I.; Jakob, C.; Dahl, A.; et al. The histone demethylase UTX regulates stem cell migration and hematopoiesis. Blood 2013, 121, 2462–2473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Wang, C.; Lee, J.E.; Cho, Y.W.; Xiao, Y.; Jin, Q.; Liu, C.; Ge, K. UTX regulates mesoderm differentiation of embryonic stem cells independent of H3K27 demethylase activity. Proc. Natl. Acad. Sci. USA 2012, 109, 15324–15329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Bogershausen, N.; Gatinois, V.; Riehmer, V.; Kayserili, H.; Becker, J.; Thoenes, M.; Simsek-Kiper, P.O.; Barat-Houari, M.; Elcioglu, N.H.; Wieczorek, D.; et al. Mutation Update for Kabuki Syndrome Genes KMT2D and KDM6A and Further Delineation of X-Linked Kabuki Syndrome Subtype 2. Hum. Mutat. 2016, 37, 847–864. [Google Scholar] [CrossRef] [Green Version]
  175. Guo, Z.; Liu, F.; Li, H.J. Novel KDM6A splice-site mutation in kabuki syndrome with congenital hydrocephalus: A case report. BMC Med. Genet. 2018, 19, 206. [Google Scholar] [CrossRef] [PubMed]
  176. Banka, S.; Lederer, D.; Benoit, V.; Jenkins, E.; Howard, E.; Bunstone, S.; Kerr, B.; McKee, S.; Lloyd, I.C.; Shears, D.; et al. Novel KDM6A (UTX) mutations and a clinical and molecular review of the X-linked Kabuki syndrome (KS2). Clin. Genet. 2015, 87, 252–258. [Google Scholar] [CrossRef]
  177. Micale, L.; Augello, B.; Maffeo, C.; Selicorni, A.; Zucchetti, F.; Fusco, C.; De Nittis, P.; Pellico, M.T.; Mandriani, B.; Fischetto, R.; et al. Molecular analysis, pathogenic mechanisms, and readthrough therapy on a large cohort of Kabuki syndrome patients. Hum. Mutat. 2014, 35, 841–850. [Google Scholar] [CrossRef]
  178. Van Laarhoven, P.M.; Neitzel, L.R.; Quintana, A.M.; Geiger, E.A.; Zackai, E.H.; Clouthier, D.E.; Artinger, K.B.; Ming, J.E.; Shaikh, T.H. Kabuki syndrome genes KMT2D and KDM6A: Functional analyses demonstrate critical roles in craniofacial, heart and brain development. Hum. Mol. Genet. 2015, 24, 4443–4453. [Google Scholar] [CrossRef] [Green Version]
  179. Miyake, N.; Koshimizu, E.; Okamoto, N.; Mizuno, S.; Ogata, T.; Nagai, T.; Kosho, T.; Ohashi, H.; Kato, M.; Sasaki, G.; et al. MLL2 and KDM6A mutations in patients with Kabuki syndrome. Am. J. Med. Genet. A 2013, 161A, 2234–2243. [Google Scholar] [CrossRef]
  180. Miyake, N.; Mizuno, S.; Okamoto, N.; Ohashi, H.; Shiina, M.; Ogata, K.; Tsurusaki, Y.; Nakashima, M.; Saitsu, H.; Niikawa, N.; et al. KDM6A point mutations cause Kabuki syndrome. Hum. Mutat. 2013, 34, 108–110. [Google Scholar] [CrossRef]
  181. Lederer, D.; Shears, D.; Benoit, V.; Verellen-Dumoulin, C.; Maystadt, I. A three generation X-linked family with Kabuki syndrome phenotype and a frameshift mutation in KDM6A. Am. J. Med. Genet. A 2014, 164A, 1289–1292. [Google Scholar] [CrossRef] [PubMed]
  182. Frans, G.; Meyts, I.; Devriendt, K.; Liston, A.; Vermeulen, F.; Bossuyt, X. Mild humoral immunodeficiency in a patient with X-linked Kabuki syndrome. Am. J. Med. Genet. A 2016, 170, 801–803. [Google Scholar] [CrossRef] [PubMed]
  183. Cheon, C.K.; Sohn, Y.B.; Ko, J.M.; Lee, Y.J.; Song, J.S.; Moon, J.W.; Yang, B.K.; Ha, I.S.; Bae, E.J.; Jin, H.S.; et al. Identification of KMT2D and KDM6A mutations by exome sequencing in Korean patients with Kabuki syndrome. J. Hum. Genet. 2014, 59, 321–325. [Google Scholar] [CrossRef] [PubMed]
  184. Masui, D.; Fukahori, S.; Mizuochi, T.; Watanabe, Y.; Fukui, K.; Ishii, S.; Saikusa, N.; Hashizume, N.; Higashidate, N.; Sakamoto, S.; et al. Cystic biliary atresia with paucity of bile ducts and gene mutation in KDM6A: A case report. Surg. Case Rep. 2019, 5, 132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. van Haaften, G.; Dalgliesh, G.L.; Davies, H.; Chen, L.; Bignell, G.; Greenman, C.; Edkins, S.; Hardy, C.; O’Meara, S.; Teague, J.; et al. Somatic mutations of the histone H3K27 demethylase gene UTX in human cancer. Nat. Genet. 2009, 41, 521–523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Ohtani, K.; Zhao, C.; Dobreva, G.; Manavski, Y.; Kluge, B.; Braun, T.; Rieger, M.A.; Zeiher, A.M.; Dimmeler, S. Jmjd3 controls mesodermal and cardiovascular differentiation of embryonic stem cells. Circ. Res. 2013, 113, 856–862. [Google Scholar] [CrossRef] [PubMed]
  187. Burgold, T.; Voituron, N.; Caganova, M.; Tripathi, P.P.; Menuet, C.; Tusi, B.K.; Spreafico, F.; Bevengut, M.; Gestreau, C.; Buontempo, S.; et al. The H3K27 demethylase JMJD3 is required for maintenance of the embryonic respiratory neuronal network, neonatal breathing, and survival. Cell Rep. 2012, 2, 1244–1258. [Google Scholar] [CrossRef] [Green Version]
  188. Satoh, T.; Takeuchi, O.; Vandenbon, A.; Yasuda, K.; Tanaka, Y.; Kumagai, Y.; Miyake, T.; Matsushita, K.; Okazaki, T.; Saitoh, T.; et al. The Jmjd3-Irf4 axis regulates M2 macrophage polarization and host responses against helminth infection. Nat. Immunol. 2010, 11, 936–944. [Google Scholar] [CrossRef]
  189. Li, Q.; Wang, H.Y.; Chepelev, I.; Zhu, Q.; Wei, G.; Zhao, K.; Wang, R.F. Stage-dependent and locus-specific role of histone demethylase Jumonji D3 (JMJD3) in the embryonic stages of lung development. PLoS Genet. 2014, 10, e1004524. [Google Scholar] [CrossRef] [Green Version]
  190. Burchfield, J.S.; Li, Q.; Wang, H.Y.; Wang, R.F. JMJD3 as an epigenetic regulator in development and disease. Int. J. Biochem. Cell Biol. 2015, 67, 148–157. [Google Scholar] [CrossRef]
  191. Zhang, F.; Xu, L.; Xu, L.; Xu, Q.; Karsenty, G.; Chen, C.D. Histone demethylase JMJD3 is required for osteoblast differentiation in mice. Sci. Rep. 2015, 5, 13418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Zhang, F.; Xu, L.; Xu, L.; Xu, Q.; Li, D.; Yang, Y.; Karsenty, G.; Chen, C.D. JMJD3 promotes chondrocyte proliferation and hypertrophy during endochondral bone formation in mice. J. Mol. Cell Biol. 2015, 7, 23–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Yavarna, T.; Al-Dewik, N.; Al-Mureikhi, M.; Ali, R.; Al-Mesaifri, F.; Mahmoud, L.; Shahbeck, N.; Lakhani, S.; AlMulla, M.; Nawaz, Z.; et al. High diagnostic yield of clinical exome sequencing in Middle Eastern patients with Mendelian disorders. Hum. Genet. 2015, 134, 967–980. [Google Scholar] [CrossRef] [PubMed]
  194. Liakath-Ali, K.; Vancollie, V.E.; Heath, E.; Smedley, D.P.; Estabel, J.; Sunter, D.; Ditommaso, T.; White, J.K.; Ramirez-Solis, R.; Smyth, I.; et al. Novel skin phenotypes revealed by a genome-wide mouse reverse genetic screen. Nat. Commun. 2014, 5, 3540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Chen, X.; Wang, S.; Zhou, Y.; Han, Y.; Li, S.; Xu, Q.; Xu, L.; Zhu, Z.; Deng, Y.; Yu, L.; et al. Phf8 histone demethylase deficiency causes cognitive impairments through the mTOR pathway. Nat. Commun. 2018, 9, 114. [Google Scholar] [CrossRef] [Green Version]
  196. Abidi, F.; Miano, M.; Murray, J.; Schwartz, C. A novel mutation in the PHF8 gene is associated with X-linked mental retardation with cleft lip/cleft palate. Clin. Genet. 2007, 72, 19–22. [Google Scholar] [CrossRef]
  197. Laumonnier, F.; Holbert, S.; Ronce, N.; Faravelli, F.; Lenzner, S.; Schwartz, C.E.; Lespinasse, J.; Van Esch, H.; Lacombe, D.; Goizet, C.; et al. Mutations in PHF8 are associated with X linked mental retardation and cleft lip/cleft palate. J. Med. Genet. 2005, 42, 780–786. [Google Scholar] [CrossRef] [Green Version]
  198. Koivisto, A.M.; Ala-Mello, S.; Lemmela, S.; Komu, H.A.; Rautio, J.; Jarvela, I. Screening of mutations in the PHF8 gene and identification of a novel mutation in a Finnish family with XLMR and cleft lip/cleft palate. Clin. Genet. 2007, 72, 145–149. [Google Scholar] [CrossRef]
  199. Nava, C.; Lamari, F.; Heron, D.; Mignot, C.; Rastetter, A.; Keren, B.; Cohen, D.; Faudet, A.; Bouteiller, D.; Gilleron, M.; et al. Analysis of the chromosome X exome in patients with autism spectrum disorders identified novel candidate genes, including TMLHE. Transl. Psychiatry 2012, 2, e179. [Google Scholar] [CrossRef] [Green Version]
  200. De Wolf, V.; Crepel, A.; Schuit, F.; van Lommel, L.; Ceulemans, B.; Steyaert, J.; Seuntjens, E.; Peeters, H.; Devriendt, K. A complex Xp11.22 deletion in a patient with syndromic autism: Exploration of FAM120C as a positional candidate gene for autism. Am. J. Med. Genet. A 2014, 164A, 3035–3041. [Google Scholar] [CrossRef]
  201. Oh, S.; Janknecht, R. Histone demethylase JMJD5 is essential for embryonic development. Biochem. Biophys. Res. Commun. 2012, 420, 61–65. [Google Scholar] [CrossRef] [PubMed]
  202. Ishimura, A.; Minehata, K.; Terashima, M.; Kondoh, G.; Hara, T.; Suzuki, T. Jmjd5, an H3K36me2 histone demethylase, modulates embryonic cell proliferation through the regulation of Cdkn1a expression. Development 2012, 139, 749–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Choudhry, H.; Harris, A.L. Advances in Hypoxia-Inducible Factor Biology. Cell Metab. 2018, 27, 281–298. [Google Scholar] [CrossRef] [PubMed]
  204. Schodel, J.; Ratcliffe, P.J. Mechanisms of hypoxia signalling: New implications for nephrology. Nat. Rev. Nephrol. 2019, 15, 641–659. [Google Scholar] [CrossRef] [PubMed]
  205. Ortiz-Barahona, A.; Villar, D.; Pescador, N.; Amigo, J.; del Peso, L. Genome-wide identification of hypoxia-inducible factor binding sites and target genes by a probabilistic model integrating transcription-profiling data and in silico binding site prediction. Nucleic Acids Res. 2010, 38, 2332–2345. [Google Scholar] [CrossRef]
  206. Benita, Y.; Kikuchi, H.; Smith, A.D.; Zhang, M.Q.; Chung, D.C.; Xavier, R.J. An integrative genomics approach identifies Hypoxia Inducible Factor-1 (HIF-1)-target genes that form the core response to hypoxia. Nucleic Acids Res. 2009, 37, 4587–4602. [Google Scholar] [CrossRef] [Green Version]
  207. Dengler, V.L.; Galbraith, M.; Espinosa, J.M. Transcriptional regulation by hypoxia inducible factors. Crit. Rev. Biochem. Mol. Biol. 2014, 49, 1–15. [Google Scholar] [CrossRef] [Green Version]
  208. Schodel, J.; Oikonomopoulos, S.; Ragoussis, J.; Pugh, C.W.; Ratcliffe, P.J.; Mole, D.R. High-resolution genome-wide mapping of HIF-binding sites by ChIP-seq. Blood 2011, 117, e207–e217. [Google Scholar] [CrossRef] [Green Version]
  209. Platt, J.L.; Salama, R.; Smythies, J.; Choudhry, H.; Davies, J.O.; Hughes, J.R.; Ratcliffe, P.J.; Mole, D.R. Capture-C reveals preformed chromatin interactions between HIF-binding sites and distant promoters. EMBO Rep. 2016, 17, 1410–1421. [Google Scholar] [CrossRef] [Green Version]
  210. Orlando, I.M.C.; Lafleur, V.N.; Storti, F.; Spielmann, P.; Crowther, L.; Santambrogio, S.; Schodel, J.; Hoogewijs, D.; Mole, D.R.; Wenger, R.H. Distal and proximal hypoxia response elements cooperate to regulate organ-specific erythropoietin gene expression. Haematologica 2019. [Google Scholar] [CrossRef] [Green Version]
  211. Kenneth, N.S.; Rocha, S. Regulation of gene expression by hypoxia. Biochem. J. 2008, 414, 19–29. [Google Scholar] [CrossRef] [Green Version]
  212. Batie, M.; Del Peso, L.; Rocha, S. Hypoxia and Chromatin: A Focus on Transcriptional Repression Mechanisms. Biomedicines 2018, 6, 47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Metzen, E.; Stiehl, D.P.; Doege, K.; Marxsen, J.H.; Hellwig-Burgel, T.; Jelkmann, W. Regulation of the prolyl hydroxylase domain protein 2 (phd2/egln-1) gene: Identification of a functional hypoxia-responsive element. Biochem. J. 2005, 387, 711–717. [Google Scholar] [CrossRef] [Green Version]
  214. Pescador, N.; Cuevas, Y.; Naranjo, S.; Alcaide, M.; Villar, D.; Landazuri, M.O.; Del Peso, L. Identification of a functional hypoxia-responsive element that regulates the expression of the egl nine homologue 3 (egln3/phd3) gene. Biochem. J. 2005, 390, 189–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Shmakova, A.; Batie, M.; Druker, J.; Rocha, S. Chromatin and oxygen sensing in the context of JmjC histone demethylases. Biochem. J. 2014, 462, 385–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Wu, M.Z.; Chen, S.F.; Nieh, S.; Benner, C.; Ger, L.P.; Jan, C.I.; Ma, L.; Chen, C.H.; Hishida, T.; Chang, H.T.; et al. Hypoxia Drives Breast Tumor Malignancy through a TET-TNFalpha-p38-MAPK Signaling Axis. Cancer Res. 2015, 75, 3912–3924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Cao, J.Z.; Liu, H.; Wickrema, A.; Godley, L.A. HIF-1 directly induces TET3 expression to enhance 5-hmC density and induce erythroid gene expression in hypoxia. Blood Adv. 2020, 4, 3053–3062. [Google Scholar] [CrossRef] [PubMed]
  218. Hu, R.; Jin, H.; Zhou, S.; Yang, P.; Li, X. Proteomic analysis of hypoxia-induced responses in the syncytialization of human placental cell line BeWo. Placenta 2007, 28, 399–407. [Google Scholar] [CrossRef] [PubMed]
  219. Sharma, N.K.; Sethy, N.K.; Bhargava, K. Comparative proteome analysis reveals differential regulation of glycolytic and antioxidant enzymes in cortex and hippocampus exposed to short-term hypobaric hypoxia. J. Proteom. 2013, 79, 277–298. [Google Scholar] [CrossRef]
  220. Hoang, V.M.; Foulk, R.; Clauser, K.; Burlingame, A.; Gibson, B.W.; Fisher, S.J. Functional proteomics: Examining the effects of hypoxia on the cytotrophoblast protein repertoire. Biochemistry 2001, 40, 4077–4086. [Google Scholar] [CrossRef]
  221. Li, Q.; Luo, T.; Lu, W.; Yi, X.; Zhao, Z.; Liu, J. Proteomic analysis of human periodontal ligament cells under hypoxia. Proteome Sci. 2019, 17, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Dou, L.; Yan, Q.; Liang, P.; Zhou, P.; Zhang, Y.; Ji, P. iTRAQ-Based Proteomic Analysis Exploring the Influence of Hypoxia on the Proteome of Dental Pulp Stem Cells under 3D Culture. Proteomics 2018, 18. [Google Scholar] [CrossRef] [PubMed]
  223. Dutta, B.; Ren, Y.; Hao, P.; Sim, K.H.; Cheow, E.; Adav, S.; Tam, J.P.; Sze, S.K. Profiling of the Chromatin-associated Proteome Identifies HP1BP3 as a Novel Regulator of Cell Cycle Progression. Mol. Cell. Proteom. 2014, 13, 2183–2197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Gao, Y.; Dasgupta, C.; Huang, L.; Song, R.; Zhang, Z.; Zhang, L. Multi-Omics Integration Reveals Short and Long-Term Effects of Gestational Hypoxia on the Heart Development. Cells 2019, 8, 1608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Shakib, K.; Norman, J.T.; Fine, L.G.; Brown, L.R.; Godovac-Zimmermann, J. Proteomics profiling of nuclear proteins for kidney fibroblasts suggests hypoxia, meiosis, and cancer may meet in the nucleus. Proteomics 2005, 5, 2819–2838. [Google Scholar] [CrossRef] [PubMed]
  226. Vodisch, M.; Scherlach, K.; Winkler, R.; Hertweck, C.; Braun, H.P.; Roth, M.; Haas, H.; Werner, E.R.; Brakhage, A.A.; Kniemeyer, O. Analysis of the Aspergillus fumigatus proteome reveals metabolic changes and the activation of the pseurotin A biosynthesis gene cluster in response to hypoxia. J. Proteome Res. 2011, 10, 2508–2524. [Google Scholar] [CrossRef] [PubMed]
  227. Janker, L.; Mayer, R.L.; Bileck, A.; Kreutz, D.; Mader, J.C.; Utpatel, K.; Heudobler, D.; Agis, H.; Gerner, C.; Slany, A. Metabolic, Anti-apoptotic and Immune Evasion Strategies of Primary Human Myeloma Cells Indicate Adaptations to Hypoxia. Mol. Cell. Proteom. 2019, 18, 936–953. [Google Scholar] [CrossRef] [PubMed]
  228. Kouzarides, T. Chromatin modifications and their function. Cell 2007, 128, 693–705. [Google Scholar] [CrossRef] [Green Version]
  229. Hsu, K.F.; Wilkins, S.E.; Hopkinson, R.J.; Sekirnik, R.; Flashman, E.; Kawamura, A.; McCullagh, J.S.O.; Walport, L.J.; Schofield, C.J. Hypoxia and hypoxia mimetics differentially modulate histone post-translational modifications. Epigenetics 2020. [Google Scholar] [CrossRef]
  230. Batie, M.; Frost, J.; Frost, M.; Wilson, J.W.; Schofield, P.; Rocha, S. Hypoxia induces rapid changes to histone methylation and reprograms chromatin. Science 2019, 363, 1222–1226. [Google Scholar] [CrossRef]
  231. Prickaerts, P.; Adriaens, M.E.; Beucken, T.V.; Koch, E.; Dubois, L.; Dahlmans, V.E.; Gits, C.; Evelo, C.T.; Chan-Seng-Yue, M.; Wouters, B.G.; et al. Hypoxia increases genome-wide bivalent epigenetic marking by specific gain of H3K27me3. Epigenet. Chromatin 2016, 9, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Lee, H.Y.; Yang, E.G.; Park, H. Hypoxia enhances the expression of prostate-specific antigen by modifying the quantity and catalytic activity of Jumonji C domain-containing histone demethylases. Carcinogenesis 2013, 34, 2706–2715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Dobrynin, G.; McAllister, T.E.; Leszczynska, K.B.; Ramachandran, S.; Krieg, A.J.; Kawamura, A.; Hammond, E.M. KDM4A regulates HIF-1 levels through H3K9me3. Sci. Rep. 2017, 7, 11094. [Google Scholar] [CrossRef] [PubMed]
  234. Luo, W.; Chang, R.; Zhong, J.; Pandey, A.; Semenza, G.L. Histone demethylase JMJD2C is a coactivator for hypoxia-inducible factor 1 that is required for breast cancer progression. Proc. Natl. Acad. Sci. USA 2012, 109, E3367–E3376. [Google Scholar] [CrossRef] [Green Version]
  235. Krieg, A.J.; Rankin, E.B.; Chan, D.; Razorenova, O.; Fernandez, S.; Giaccia, A.J. Regulation of the histone demethylase JMJD1A by hypoxia-inducible factor 1 alpha enhances hypoxic gene expression and tumor growth. Mol. Cell. Biol. 2010, 30, 344–353. [Google Scholar] [CrossRef] [Green Version]
  236. Wan, W.; Peng, K.; Li, M.; Qin, L.; Tong, Z.; Yan, J.; Shen, B.; Yu, C. Histone demethylase JMJD1A promotes urinary bladder cancer progression by enhancing glycolysis through coactivation of hypoxia inducible factor 1alpha. Oncogene 2017, 36, 3868–3877. [Google Scholar] [CrossRef]
  237. Qian, X.; Li, X.; Shi, Z.; Bai, X.; Xia, Y.; Zheng, Y.; Xu, D.; Chen, F.; You, Y.; Fang, J.; et al. KDM3A Senses Oxygen Availability to Regulate PGC-1alpha-Mediated Mitochondrial Biogenesis. Mol. Cell. 2019, 76, 885–895 e887. [Google Scholar] [CrossRef]
  238. Wu, X.; Zhang, Y. TET-mediated active DNA demethylation: Mechanism, function and beyond. Nat. Rev. Genet. 2017, 18, 517–534. [Google Scholar] [CrossRef]
  239. Thienpont, B.; Steinbacher, J.; Zhao, H.; D’Anna, F.; Kuchnio, A.; Ploumakis, A.; Ghesquiere, B.; Van Dyck, L.; Boeckx, B.; Schoonjans, L.; et al. Tumour hypoxia causes DNA hypermethylation by reducing TET activity. Nature 2016, 537, 63–68. [Google Scholar] [CrossRef]
  240. Shahrzad, S.; Bertrand, K.; Minhas, K.; Coomber, B.L. Induction of DNA hypomethylation by tumor hypoxia. Epigenetics 2007, 2, 119–125. [Google Scholar] [CrossRef] [Green Version]
  241. D’Anna, F.; Van Dyck, L.; Xiong, J.; Zhao, H.; Berrens, R.V.; Qian, J.; Bieniasz-Krzywiec, P.; Chandra, V.; Schoonjans, L.; Matthews, J.; et al. DNA methylation repels binding of hypoxia-inducible transcription factors to maintain tumor immunotolerance. Genome Biol. 2020, 21, 182. [Google Scholar] [CrossRef] [PubMed]
  242. Camuzi, D.; de Amorim, I.S.S.; Ribeiro Pinto, L.F.; Oliveira Trivilin, L.; Mencalha, A.L.; Soares Lima, S.C. Regulation Is in the Air: The Relationship between Hypoxia and Epigenetics in Cancer. Cells 2019, 8, 300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Ito, S.; D’Alessio, A.C.; Taranova, O.V.; Hong, K.; Sowers, L.C.; Zhang, Y. Role of Tet proteins in 5mC to 5hmC conversion, ES-cell self-renewal and inner cell mass specification. Nature 2010, 466, 1129–1133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Tahiliani, M.; Koh, K.P.; Shen, Y.; Pastor, W.A.; Bandukwala, H.; Brudno, Y.; Agarwal, S.; Iyer, L.M.; Liu, D.R.; Aravind, L.; et al. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 2009, 324, 930–935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Brugarolas, J.; Lei, K.; Hurley, R.L.; Manning, B.D.; Reiling, J.H.; Hafen, E.; Witters, L.A.; Ellisen, L.W.; Kaelin, W.G., Jr. Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes Dev. 2004, 18, 2893–2904. [Google Scholar] [CrossRef] [Green Version]
  246. Connolly, E.; Braunstein, S.; Formenti, S.; Schneider, R.J. Hypoxia inhibits protein synthesis through a 4E-BP1 and elongation factor 2 kinase pathway controlled by mTOR and uncoupled in breast cancer cells. Mol. Cell. Biol. 2006, 26, 3955–3965. [Google Scholar] [CrossRef] [Green Version]
  247. Sonenberg, N.; Hinnebusch, A.G. Regulation of translation initiation in eukaryotes: Mechanisms and biological targets. Cell 2009, 136, 731–745. [Google Scholar] [CrossRef] [Green Version]
  248. Koumenis, C.; Naczki, C.; Koritzinsky, M.; Rastani, S.; Diehl, A.; Sonenberg, N.; Koromilas, A.; Wouters, B.G. Regulation of protein synthesis by hypoxia via activation of the endoplasmic reticulum kinase PERK and phosphorylation of the translation initiation factor eIF2alpha. Mol. Cell. Biol. 2002, 22, 7405–7416. [Google Scholar] [CrossRef] [Green Version]
  249. Donnelly, N.; Gorman, A.M.; Gupta, S.; Samali, A. The eIF2alpha kinases: Their structures and functions. Cell. Mol. Life Sci. 2013, 70, 3493–3511. [Google Scholar] [CrossRef]
  250. Ryazanov, A.G.; Davydova, E.K. Mechanism of elongation factor 2 (EF-2) inactivation upon phosphorylation. Phosphorylated EF-2 is unable to catalyze translocation. FEBS Lett. 1989, 251, 187–190. [Google Scholar] [CrossRef] [Green Version]
  251. Browne, G.J.; Proud, C.G. A novel mTOR-regulated phosphorylation site in elongation factor 2 kinase modulates the activity of the kinase and its binding to calmodulin. Mol. Cell. Biol. 2004, 24, 2986–2997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Liu, L.; Cash, T.P.; Jones, R.G.; Keith, B.; Thompson, C.B.; Simon, M.C. Hypoxia-induced energy stress regulates mRNA translation and cell growth. Mol. Cell. 2006, 21, 521–531. [Google Scholar] [CrossRef] [Green Version]
  253. Moore, C.E.; Mikolajek, H.; Regufe da Mota, S.; Wang, X.; Kenney, J.W.; Werner, J.M.; Proud, C.G. Elongation Factor 2 Kinase Is Regulated by Proline Hydroxylation and Protects Cells during Hypoxia. Mol. Cell. Biol. 2015, 35, 1788–1804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Pollard, P.J.; Loenarz, C.; Mole, D.R.; McDonough, M.A.; Gleadle, J.M.; Schofield, C.J.; Ratcliffe, P.J. Regulation of Jumonji-domain-containing histone demethylases by hypoxia-inducible factor (HIF)-1alpha. Biochem. J. 2008, 416, 387–394. [Google Scholar] [CrossRef] [PubMed]
  255. Kato, M.; Araiso, Y.; Noma, A.; Nagao, A.; Suzuki, T.; Ishitani, R.; Nureki, O. Crystal structure of a novel JmjC-domain-containing protein, TYW5, involved in tRNA modification. Nucleic Acids Res. 2011, 39, 1576–1585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Ge, W.; Wolf, A.; Feng, T.; Ho, C.H.; Sekirnik, R.; Zayer, A.; Granatino, N.; Cockman, M.E.; Loenarz, C.; Loik, N.D.; et al. Oxygenase-catalyzed ribosome hydroxylation occurs in prokaryotes and humans. Nat. Chem. Biol. 2012, 8, 960–962. [Google Scholar] [CrossRef] [PubMed]
  257. Loenarz, C.; Sekirnik, R.; Thalhammer, A.; Ge, W.; Spivakovsky, E.; Mackeen, M.M.; McDonough, M.A.; Cockman, M.E.; Kessler, B.M.; Ratcliffe, P.J.; et al. Hydroxylation of the eukaryotic ribosomal decoding center affects translational accuracy. Proc. Natl. Acad. Sci. USA 2014, 111, 4019–4024. [Google Scholar] [CrossRef] [Green Version]
  258. Singleton, R.S.; Liu-Yi, P.; Formenti, F.; Ge, W.; Sekirnik, R.; Fischer, R.; Adam, J.; Pollard, P.J.; Wolf, A.; Thalhammer, A.; et al. OGFOD1 catalyzes prolyl hydroxylation of RPS23 and is involved in translation control and stress granule formation. Proc. Natl. Acad. Sci. USA 2014, 111, 4031–4036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Fu, D.; Brophy, J.A.; Chan, C.T.; Atmore, K.A.; Begley, U.; Paules, R.S.; Dedon, P.C.; Begley, T.J.; Samson, L.D. Human AlkB homolog ABH8 Is a tRNA methyltransferase required for wobble uridine modification and DNA damage survival. Mol. Cell. Biol. 2010, 30, 2449–2459. [Google Scholar] [CrossRef] [Green Version]
  260. Noma, A.; Ishitani, R.; Kato, M.; Nagao, A.; Nureki, O.; Suzuki, T. Expanding role of the jumonji C domain as an RNA hydroxylase. J. Biol. Chem. 2010, 285, 34503–34507. [Google Scholar] [CrossRef] [Green Version]
  261. Feng, T.; Yamamoto, A.; Wilkins, S.E.; Sokolova, E.; Yates, L.A.; Munzel, M.; Singh, P.; Hopkinson, R.J.; Fischer, R.; Cockman, M.E.; et al. Optimal translational termination requires C4 lysyl hydroxylation of eRF1. Mol. Cell. 2014, 53, 645–654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Zhou, T.; Erber, L.; Liu, B.; Gao, Y.; Ruan, H.B.; Chen, Y. Proteomic analysis reveals diverse proline hydroxylation-mediated oxygen-sensing cellular pathways in cancer cells. Oncotarget 2016, 7, 79154–79169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Cockman, M.E.; Webb, J.D.; Kramer, H.B.; Kessler, B.M.; Ratcliffe, P.J. Proteomics-based identification of novel factor inhibiting hypoxia-inducible factor (FIH) substrates indicates widespread asparaginyl hydroxylation of ankyrin repeat domain-containing proteins. Mol. Cell. Proteom. 2009, 8, 535–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Malec, V.; Coulson, J.M.; Urbe, S.; Clague, M.J. Combined Analyses of the VHL and Hypoxia Signaling Axes in an Isogenic Pairing of Renal Clear Cell Carcinoma Cells. J. Proteome Res. 2015, 14, 5263–5272. [Google Scholar] [CrossRef] [PubMed]
  265. Silverman-Gavrila, L.B.; Lu, T.Z.; Prashad, R.C.; Nejatbakhsh, N.; Charlton, M.P.; Feng, Z.P. Neural phosphoproteomics of a chronic hypoxia model--Lymnaea stagnalis. Neuroscience 2009, 161, 621–634. [Google Scholar] [CrossRef]
  266. Chachami, G.; Stankovic-Valentin, N.; Karagiota, A.; Basagianni, A.; Plessmann, U.; Urlaub, H.; Melchior, F.; Simos, G. Hypoxia-induced Changes in SUMO Conjugation Affect Transcriptional Regulation Under Low Oxygen. Mol. Cell. Proteom. 2019, 18, 1197–1209. [Google Scholar] [CrossRef] [Green Version]
  267. Dhillon, R.S.; Richards, J.G. Hypoxia induces selective modifications to the acetylome in the brain of zebrafish (Danio rerio). Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2018, 224, 79–87. [Google Scholar] [CrossRef]
  268. Arriagada, C.; Silva, P.; Torres, V.A. Role of glycosylation in hypoxia-driven cell migration and invasion. Cell Adhes. Migr. 2019, 13, 13–22. [Google Scholar] [CrossRef] [Green Version]
  269. Peinado, M.A.; Hernandez, R.; Peragon, J.; Ovelleiro, D.; Pedrosa, J.A.; Blanco, S. Proteomic characterization of nitrated cell targets after hypobaric hypoxia and reoxygenation in rat brain. J. Proteom. 2014, 109, 309–321. [Google Scholar] [CrossRef] [PubMed]
  270. Chen, S.C.; Huang, B.; Liu, Y.C.; Shyu, K.G.; Lin, P.Y.; Wang, D.L. Acute hypoxia enhances proteins’ S-nitrosylation in endothelial cells. Biochem. Biophys. Res. Commun. 2008, 377, 1274–1278. [Google Scholar] [CrossRef]
  271. Kumar, G.K.; Prabhakar, N.R. Post-translational modification of proteins during intermittent hypoxia. Respir. Physiol. Neurobiol. 2008, 164, 272–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Moser, S.C.; Bensaddek, D.; Ortmann, B.; Maure, J.F.; Mudie, S.; Blow, J.J.; Lamond, A.I.; Swedlow, J.R.; Rocha, S. PHD1 links cell-cycle progression to oxygen sensing through hydroxylation of the centrosomal protein Cep192. Dev. Cell 2013, 26, 381–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Zheng, X.; Zhai, B.; Koivunen, P.; Shin, S.J.; Lu, G.; Liu, J.; Geisen, C.; Chakraborty, A.A.; Moslehi, J.J.; Smalley, D.M.; et al. Prolyl hydroxylation by EglN2 destabilizes FOXO3a by blocking its interaction with the USP9x deubiquitinase. Genes Dev. 2014, 28, 1429–1444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Zurlo, G.; Liu, X.; Takada, M.; Fan, C.; Simon, J.M.; Ptacek, T.S.; Rodriguez, J.; von Kriegsheim, A.; Liu, J.; Locasale, J.W.; et al. Prolyl hydroxylase substrate adenylosuccinate lyase is an oncogenic driver in triple negative breast cancer. Nat. Commun. 2019, 10, 5177. [Google Scholar] [CrossRef] [PubMed]
  275. Luo, W.; Lin, B.; Wang, Y.; Zhong, J.; O’Meally, R.; Cole, R.N.; Pandey, A.; Levchenko, A.; Semenza, G.L. PHD3-mediated prolyl hydroxylation of nonmuscle actin impairs polymerization and cell motility. Mol. Biol. Cell. 2014, 25, 2788–2796. [Google Scholar] [CrossRef] [Green Version]
  276. Hu, L.; Xie, H.; Liu, X.; Potjewyd, F.; James, L.I.; Wilkerson, E.M.; Herring, L.E.; Xie, L.; Chen, X.; Cabrera, J.C.; et al. TBK1 Is a Synthetic Lethal Target in Cancer with VHL Loss. Cancer Discov. 2020, 10, 460–475. [Google Scholar] [CrossRef]
  277. Liu, X.; Simon, J.M.; Xie, H.; Hu, L.; Wang, J.; Zurlo, G.; Fan, C.; Ptacek, T.S.; Herring, L.; Tan, X.; et al. Genome-wide Screening Identifies SFMBT1 as an Oncogenic Driver in Cancer with VHL Loss. Mol. Cell. 2020, 77, 1294–1306 e1295. [Google Scholar] [CrossRef]
  278. Guo, J.; Chakraborty, A.A.; Liu, P.; Gan, W.; Zheng, X.; Inuzuka, H.; Wang, B.; Zhang, J.; Zhang, L.; Yuan, M.; et al. pVHL suppresses kinase activity of Akt in a proline-hydroxylation-dependent manner. Science 2016, 353, 929–932. [Google Scholar] [CrossRef] [Green Version]
  279. Cockman, M.E.; Lippl, K.; Tian, Y.M.; Pegg, H.B.; Figg, W.D.J.; Abboud, M.I.; Heilig, R.; Fischer, R.; Myllyharju, J.; Schofield, C.J.; et al. Lack of activity of recombinant HIF prolyl hydroxylases (PHDs) on reported non-HIF substrates. Elife 2019, 8. [Google Scholar] [CrossRef]
  280. Lee, S.B.; Ko, A.; Oh, Y.T.; Shi, P.; D’Angelo, F.; Frangaj, B.; Koller, A.; Chen, E.I.; Cardozo, T.; Iavarone, A.; et al. Proline Hydroxylation Primes Protein Kinases for Autophosphorylation and Activation. Mol. Cell. 2020, 79, 376–389.e378. [Google Scholar] [CrossRef]
  281. Arsenault, P.R.; Heaton-Johnson, K.J.; Li, L.S.; Song, D.; Ferreira, V.S.; Patel, N.; Master, S.R.; Lee, F.S. Identification of prolyl hydroxylation modifications in mammalian cell proteins. Proteomics 2015, 15, 1259–1267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Chang, B.; Chen, Y.; Zhao, Y.; Bruick, R.K. JMJD6 is a histone arginine demethylase. Science 2007, 318, 444–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Islam, M.S.; McDonough, M.A.; Chowdhury, R.; Gault, J.; Khan, A.; Pires, E.; Schofield, C.J. Biochemical and structural investigations clarify the substrate selectivity of the 2-oxoglutarate oxygenase JMJD6. J. Biol. Chem. 2019, 294, 11637–11652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Alahari, S.; Post, M.; Caniggia, I. Jumonji Domain Containing Protein 6: A Novel Oxygen Sensor in the Human Placenta. Endocrinology 2015, 156, 3012–3025. [Google Scholar] [CrossRef]
  285. Tsai, W.C.; Reineke, L.C.; Jain, A.; Jung, S.Y.; Lloyd, R.E. Histone arginine demethylase JMJD6 is linked to stress granule assembly through demethylation of the stress granule-nucleating protein G3BP1. J. Biol. Chem. 2017, 292, 18886–18896. [Google Scholar] [CrossRef] [Green Version]
  286. Tsai, W.C.; Gayatri, S.; Reineke, L.C.; Sbardella, G.; Bedford, M.T.; Lloyd, R.E. Arginine Demethylation of G3BP1 Promotes Stress Granule Assembly. J. Biol. Chem. 2016, 291, 22671–22685. [Google Scholar] [CrossRef] [Green Version]
  287. Webby, C.J.; Wolf, A.; Gromak, N.; Dreger, M.; Kramer, H.; Kessler, B.; Nielsen, M.L.; Schmitz, C.; Butler, D.S.; Yates, J.R., 3rd; et al. Jmjd6 catalyses lysyl-hydroxylation of U2AF65, a protein associated with RNA splicing. Science 2009, 325, 90–93. [Google Scholar] [CrossRef]
  288. Kwok, J.; O’Shea, M.; Hume, D.A.; Lengeling, A. Jmjd6, a JmjC Dioxygenase with Many Interaction Partners and Pleiotropic Functions. Front. Genet. 2017, 8, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  289. Liu, W.; Ma, Q.; Wong, K.; Li, W.; Ohgi, K.; Zhang, J.; Aggarwal, A.; Rosenfeld, M.G. Brd4 and JMJD6-associated anti-pause enhancers in regulation of transcriptional pause release. Cell 2013, 155, 1581–1595. [Google Scholar] [CrossRef] [Green Version]
  290. Wang, F.; He, L.; Huangyang, P.; Liang, J.; Si, W.; Yan, R.; Han, X.; Liu, S.; Gui, B.; Li, W.; et al. JMJD6 promotes colon carcinogenesis through negative regulation of p53 by hydroxylation. PLoS Biol. 2014, 12, e1001819. [Google Scholar] [CrossRef] [Green Version]
  291. Han, G.; Li, J.; Wang, Y.; Li, X.; Mao, H.; Liu, Y.; Chen, C.D. The hydroxylation activity of Jmjd6 is required for its homo-oligomerization. J. Cell. Biochem. 2012, 113, 1663–1670. [Google Scholar] [CrossRef]
  292. Tibrewal, N.; Liu, T.; Li, H.; Birge, R.B. Characterization of the biochemical and biophysical properties of the phosphatidylserine receptor (PS-R) gene product. Mol. Cell. Biochem. 2007, 304, 119–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Wilkins, S.E.; Islam, M.S.; Gannon, J.M.; Markolovic, S.; Hopkinson, R.J.; Ge, W.; Schofield, C.J.; Chowdhury, R. JMJD5 is a human arginyl C-3 hydroxylase. Nat. Commun. 2018, 9, 1180. [Google Scholar] [CrossRef] [PubMed]
  294. Shen, J.; Xiang, X.; Chen, L.; Wang, H.; Wu, L.; Sun, Y.; Ma, L.; Gu, X.; Liu, H.; Wang, L.; et al. JMJD5 cleaves monomethylated histone H3 N-tail under DNA damaging stress. EMBO Rep. 2017, 18, 2131–2143. [Google Scholar] [CrossRef] [PubMed]
  295. Liu, H.; Wang, C.; Lee, S.; Deng, Y.; Wither, M.; Oh, S.; Ning, F.; Dege, C.; Zhang, Q.; Liu, X.; et al. Clipping of arginine-methylated histone tails by JMJD5 and JMJD7. Proc. Natl. Acad. Sci. USA 2017, 114, E7717–E7726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Liu, H.; Wang, C.; Lee, S.; Ning, F.; Wang, Y.; Zhang, Q.; Chen, Z.; Zang, J.; Nix, J.; Dai, S.; et al. Specific Recognition of Arginine Methylated Histone Tails by JMJD5 and JMJD7. Sci. Rep. 2018, 8, 3275. [Google Scholar] [CrossRef] [PubMed]
  297. Hsia, D.A.; Tepper, C.G.; Pochampalli, M.R.; Hsia, E.Y.; Izumiya, C.; Huerta, S.B.; Wright, M.E.; Chen, H.W.; Kung, H.J.; Izumiya, Y. KDM8, a H3K36me2 histone demethylase that acts in the cyclin A1 coding region to regulate cancer cell proliferation. Proc. Natl. Acad. Sci. USA 2010, 107, 9671–9676. [Google Scholar] [CrossRef] [Green Version]
  298. Marcon, E.; Ni, Z.; Pu, S.; Turinsky, A.L.; Trimble, S.S.; Olsen, J.B.; Silverman-Gavrila, R.; Silverman-Gavrila, L.; Phanse, S.; Guo, H.; et al. Human-chromatin-related protein interactions identify a demethylase complex required for chromosome segregation. Cell Rep. 2014, 8, 297–310. [Google Scholar] [CrossRef] [Green Version]
  299. Huang, X.; Zhang, S.; Qi, H.; Wang, Z.; Chen, H.W.; Shao, J.; Shen, J. JMJD5 interacts with p53 and negatively regulates p53 function in control of cell cycle and proliferation. Biochim. Biophys. Acta 2015, 1853, 2286–2295. [Google Scholar] [CrossRef] [Green Version]
  300. Ishimura, A.; Terashima, M.; Tange, S.; Suzuki, T. Jmjd5 functions as a regulator of p53 signaling during mouse embryogenesis. Cell Tissue Res. 2016, 363, 723–733. [Google Scholar] [CrossRef]
  301. Markolovic, S.; Zhuang, Q.; Wilkins, S.E.; Eaton, C.D.; Abboud, M.I.; Katz, M.J.; McNeil, H.E.; Lesniak, R.K.; Hall, C.; Struwe, W.B.; et al. The Jumonji-C oxygenase JMJD7 catalyzes (3S)-lysyl hydroxylation of TRAFAC GTPases. Nat. Chem. Biol. 2018, 14, 688–695. [Google Scholar] [CrossRef] [Green Version]
  302. Lu, T.; Jackson, M.W.; Wang, B.; Yang, M.; Chance, M.R.; Miyagi, M.; Gudkov, A.V.; Stark, G.R. Regulation of NF-kappaB by NSD1/FBXL11-dependent reversible lysine methylation of p65. Proc. Natl. Acad. Sci. USA 2010, 107, 46–51. [Google Scholar] [CrossRef] [Green Version]
  303. Ramadoss, S.; Guo, G.; Wang, C.Y. Lysine demethylase KDM3A regulates breast cancer cell invasion and apoptosis by targeting histone and the non-histone protein p53. Oncogene 2017, 36, 47–59. [Google Scholar] [CrossRef] [Green Version]
  304. Ponnaluri, V.K.; Vavilala, D.T.; Putty, S.; Gutheil, W.G.; Mukherji, M. Identification of non-histone substrates for JMJD2A-C histone demethylases. Biochem. Biophys. Res. Commun. 2009, 390, 280–284. [Google Scholar] [CrossRef] [PubMed]
  305. Walport, L.J.; Hopkinson, R.J.; Chowdhury, R.; Schiller, R.; Ge, W.; Kawamura, A.; Schofield, C.J. Arginine demethylation is catalysed by a subset of JmjC histone lysine demethylases. Nat. Commun. 2016, 7, 11974. [Google Scholar] [CrossRef] [PubMed]
  306. Meyer, K.D.; Patil, D.P.; Zhou, J.; Zinoviev, A.; Skabkin, M.A.; Elemento, O.; Pestova, T.V.; Qian, S.B.; Jaffrey, S.R. 5’ UTR m(6)A Promotes Cap-Independent Translation. Cell 2015, 163, 999–1010. [Google Scholar] [CrossRef] [Green Version]
  307. Stenson, P.D.; Mort, M.; Ball, E.V.; Shaw, K.; Phillips, A.; Cooper, D.N. The Human Gene Mutation Database: Building a comprehensive mutation repository for clinical and molecular genetics, diagnostic testing and personalized genomic medicine. Hum. Genet. 2014, 133, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Ang, S.O.; Chen, H.; Hirota, K.; Gordeuk, V.R.; Jelinek, J.; Guan, Y.; Liu, E.; Sergueeva, A.I.; Miasnikova, G.Y.; Mole, D.; et al. Disruption of oxygen homeostasis underlies congenital Chuvash polycythemia. Nat. Genet. 2002, 32, 614–621. [Google Scholar] [CrossRef]
  309. Gordeuk, V.R.; Sergueeva, A.I.; Miasnikova, G.Y.; Okhotin, D.; Voloshin, Y.; Choyke, P.L.; Butman, J.A.; Jedlickova, K.; Prchal, J.T.; Polyakova, L.A. Congenital disorder of oxygen sensing: Association of the homozygous Chuvash polycythemia VHL mutation with thrombosis and vascular abnormalities but not tumors. Blood 2004, 103, 3924–3932. [Google Scholar] [CrossRef]
  310. Smith, T.G.; Brooks, J.T.; Balanos, G.M.; Lappin, T.R.; Layton, D.M.; Leedham, D.L.; Liu, C.; Maxwell, P.H.; McMullin, M.F.; McNamara, C.J.; et al. Mutation of von Hippel-Lindau tumour suppressor and human cardiopulmonary physiology. PLoS Med. 2006, 3, e290. [Google Scholar] [CrossRef] [Green Version]
  311. Rasmussen, K.D.; Helin, K. Role of TET enzymes in DNA methylation, development, and cancer. Genes Dev. 2016, 30, 733–750. [Google Scholar] [CrossRef] [PubMed]
  312. Delhommeau, F.; Dupont, S.; Della Valle, V.; James, C.; Trannoy, S.; Masse, A.; Kosmider, O.; Le Couedic, J.P.; Robert, F.; Alberdi, A.; et al. Mutation in TET2 in myeloid cancers. N. Engl. J. Med. 2009, 360, 2289–2301. [Google Scholar] [CrossRef] [PubMed]
  313. Lederer, D.; Grisart, B.; Digilio, M.C.; Benoit, V.; Crespin, M.; Ghariani, S.C.; Maystadt, I.; Dallapiccola, B.; Verellen-Dumoulin, C. Deletion of KDM6A, a histone demethylase interacting with MLL2, in three patients with Kabuki syndrome. Am. J. Hum. Genet. 2012, 90, 119–124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Yang, P.; Tan, H.; Xia, Y.; Yu, Q.; Wei, X.; Guo, R.; Peng, Y.; Chen, C.; Li, H.; Mei, L.; et al. De novo exonic deletion of KDM6A in a Chinese girl with Kabuki syndrome: A case report and brief literature review. Am. J. Med. Genet. A 2016, 170, 1613–1621. [Google Scholar] [CrossRef] [PubMed]
  315. Lindgren, A.M.; Hoyos, T.; Talkowski, M.E.; Hanscom, C.; Blumenthal, I.; Chiang, C.; Ernst, C.; Pereira, S.; Ordulu, Z.; Clericuzio, C.; et al. Haploinsufficiency of KDM6A is associated with severe psychomotor retardation, global growth restriction, seizures and cleft palate. Hum. Genet. 2013, 132, 537–552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Dunwoodie, S.L. The role of hypoxia in development of the Mammalian embryo. Dev Cell. 2009, 17, 755–773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Petrova, V.; Annicchiarico-Petruzzelli, M.; Melino, G.; Amelio, I. The hypoxic tumour microenvironment. Oncogenesis 2018, 7, 10. [Google Scholar] [CrossRef]
  318. Kaelin, W.G., Jr. Cancer and altered metabolism: Potential importance of hypoxia-inducible factor and 2-oxoglutarate-dependent dioxygenases. Cold Spring Harb. Symp. Quant. Biol. 2011, 76, 335–345. [Google Scholar] [CrossRef]
  319. Chang, S.; Yim, S.; Park, H. The cancer driver genes IDH1/2, JARID1C/ KDM5C, and UTX/ KDM6A: Crosstalk between histone demethylation and hypoxic reprogramming in cancer metabolism. Exp. Mol. Med. 2019, 51, 1–17. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Regulation of HIF levels and activity in normoxia and hypoxia. Under normal oxygen conditions, (A), normoxia, HIF-α is constantly hydroxylated by PHDs and FIH. PHD-mediated hydroxylation increases binding affinity with the tumour suppressor VHL, which promotes ubiquitination and degradation by the proteosome. As oxygen levels decrease, in mild hypoxia (15–1% O2) (B), PHDs are inhibited, HIF-α is stabilised, though still hydroxylated by FIH, binds to HIF-1β and is able to induce transcription of certain target genes. With further reduction in oxygen levels, in severe hypoxia (<1% O2) (C), FIH is also inhibited and HIF is able to become fully active by the recruitment of co-activators such as p300.
Figure 1. Regulation of HIF levels and activity in normoxia and hypoxia. Under normal oxygen conditions, (A), normoxia, HIF-α is constantly hydroxylated by PHDs and FIH. PHD-mediated hydroxylation increases binding affinity with the tumour suppressor VHL, which promotes ubiquitination and degradation by the proteosome. As oxygen levels decrease, in mild hypoxia (15–1% O2) (B), PHDs are inhibited, HIF-α is stabilised, though still hydroxylated by FIH, binds to HIF-1β and is able to induce transcription of certain target genes. With further reduction in oxygen levels, in severe hypoxia (<1% O2) (C), FIH is also inhibited and HIF is able to become fully active by the recruitment of co-activators such as p300.
Cancers 13 00350 g001
Figure 2. Control of gene expression by 2-ODGs in hypoxia. Hypoxia via the regulation of 2-OGDs has the potential of controlling all aspects of gene expression and protein function. Action of JmjC-histone demethylases and TETs (mediators of DNA demethylation) will impact on chromatin and DNA regulation. RNA demethylases, and several hydroxylases acts on mRNA processing, and fate. Hydroxylases also control rates of translation and ribosome activity, while JmjC-demethylases and hydroxylases can control protein function directly or indirectly by controlling other PTMs.
Figure 2. Control of gene expression by 2-ODGs in hypoxia. Hypoxia via the regulation of 2-OGDs has the potential of controlling all aspects of gene expression and protein function. Action of JmjC-histone demethylases and TETs (mediators of DNA demethylation) will impact on chromatin and DNA regulation. RNA demethylases, and several hydroxylases acts on mRNA processing, and fate. Hydroxylases also control rates of translation and ribosome activity, while JmjC-demethylases and hydroxylases can control protein function directly or indirectly by controlling other PTMs.
Cancers 13 00350 g002
Figure 3. Chromatin oxygen sensing via JmjC histone demethylases and TETs. JmjC histone demethylases and TETs (mediators DNA demethylation) are 2-OGDs. Reduced activity of these enzymes in hypoxia, due to their oxygen sensitivity, can alter the chromatin landscape and mediate hypoxia induced transcription changes. Reduced activity of KDM5A in hypoxia increases H3K4me3 at the promoters of a subset hypoxia induced genes, facilitating their transcriptional activation. Reduced activity of KDM6A in hypoxia increases H3K27me3 at the promoters of a subset of hypoxia-repressed genes and represses their transcription. Reduced TET1/2 activity in hypoxia also represses gene transcription via DNA hypermethylation at gene promoters.
Figure 3. Chromatin oxygen sensing via JmjC histone demethylases and TETs. JmjC histone demethylases and TETs (mediators DNA demethylation) are 2-OGDs. Reduced activity of these enzymes in hypoxia, due to their oxygen sensitivity, can alter the chromatin landscape and mediate hypoxia induced transcription changes. Reduced activity of KDM5A in hypoxia increases H3K4me3 at the promoters of a subset hypoxia induced genes, facilitating their transcriptional activation. Reduced activity of KDM6A in hypoxia increases H3K27me3 at the promoters of a subset of hypoxia-repressed genes and represses their transcription. Reduced TET1/2 activity in hypoxia also represses gene transcription via DNA hypermethylation at gene promoters.
Cancers 13 00350 g003
Figure 4. Hypoxia induces a global inhibition of protein translation. Global translation is mostly inhibited at initiation through mTOR inhibition of eIF4 and PERK inhibition of eIF2α, subsequently decreasing cap-dependent translation of mRNA. mTOR is inhibited through hypoxia-induced DDIT4-dependent release of TSC2 from 14-3-3 binding proteins, resulting in the inhibition of mTOR by TSC1/2 dimer. Elongation is also regulated through mTOR, as well as AMPK through its inhibition of RPS6KB1, wherein hypoxia eEF2K is not inhibited, which allows its phosphorylation and inhibition of eEF2. PHD2 can also hydroxylate eEF2K, which in normoxia causes its disassociation from calmodulin, decreasing its autophosphorylation. Termination is regulated by JMJD4-mediated hydroxylation of eRF1, which is required for termination. Selective translation of genes in hypoxia is regulated by UTR sequences, such as IRES and uORFs, which allow increased translation specifically in hypoxia. RNA binding proteins can bind to various parts of the mRNA and result in different regulatory outcomes. The ribosome can also be hydroxylated by RIOX1 and RIOX2, though it is not yet clear what role these modifications have.
Figure 4. Hypoxia induces a global inhibition of protein translation. Global translation is mostly inhibited at initiation through mTOR inhibition of eIF4 and PERK inhibition of eIF2α, subsequently decreasing cap-dependent translation of mRNA. mTOR is inhibited through hypoxia-induced DDIT4-dependent release of TSC2 from 14-3-3 binding proteins, resulting in the inhibition of mTOR by TSC1/2 dimer. Elongation is also regulated through mTOR, as well as AMPK through its inhibition of RPS6KB1, wherein hypoxia eEF2K is not inhibited, which allows its phosphorylation and inhibition of eEF2. PHD2 can also hydroxylate eEF2K, which in normoxia causes its disassociation from calmodulin, decreasing its autophosphorylation. Termination is regulated by JMJD4-mediated hydroxylation of eRF1, which is required for termination. Selective translation of genes in hypoxia is regulated by UTR sequences, such as IRES and uORFs, which allow increased translation specifically in hypoxia. RNA binding proteins can bind to various parts of the mRNA and result in different regulatory outcomes. The ribosome can also be hydroxylated by RIOX1 and RIOX2, though it is not yet clear what role these modifications have.
Cancers 13 00350 g004
Figure 5. JmjC-domain containing 2OG dioxygenases have diverse cellular functions. JMJD6 is able to promote the formation of stress granules, which reversibly pauses transcript translation through demethylation of G3BP1. JMJD6 can hydroxylate splicing regulatory (SR) proteins, resulting in differential splicing or exon choice, such as skipping the first exon with the hydroxylation of SRSF11. JMJD6 can also hydroxylate p53 repressing its activity. KDM8 is an arginine hydroxylase that can target RCCD1 for chromatin condensation, and RPS6, which affects translation. JMJD7 is a lysyl hydroxylase that can target DRG1/2, affecting their regulation of RNA. KDM2A represses NF-κB activity via demethylation of RELA. KDM3A can demethylate both PGC-1α, promoting mitochondrial biogenesis, and p53, inhibiting its proapoptotic activity. KDM3B is able to demethylate mono- or di-methylated arginine residues, which functions to activate gene transcription.
Figure 5. JmjC-domain containing 2OG dioxygenases have diverse cellular functions. JMJD6 is able to promote the formation of stress granules, which reversibly pauses transcript translation through demethylation of G3BP1. JMJD6 can hydroxylate splicing regulatory (SR) proteins, resulting in differential splicing or exon choice, such as skipping the first exon with the hydroxylation of SRSF11. JMJD6 can also hydroxylate p53 repressing its activity. KDM8 is an arginine hydroxylase that can target RCCD1 for chromatin condensation, and RPS6, which affects translation. JMJD7 is a lysyl hydroxylase that can target DRG1/2, affecting their regulation of RNA. KDM2A represses NF-κB activity via demethylation of RELA. KDM3A can demethylate both PGC-1α, promoting mitochondrial biogenesis, and p53, inhibiting its proapoptotic activity. KDM3B is able to demethylate mono- or di-methylated arginine residues, which functions to activate gene transcription.
Cancers 13 00350 g005
Table 1. 2-OGDs with reported affinities for oxygen from in vitro assays. The KM values for oxygen have been determined via a range of in vitro experiments. The KM obtained is affected by substrate, conditions, and technique, and therefore provides only a possible indication of oxygen sensing capability. We have assessed, from the reported KMs and cellular data, whether these 2-OGDs might act as oxygen sensors. * Median KM from multiple studies (individual values listed in Supplementary Table S1).
Table 1. 2-OGDs with reported affinities for oxygen from in vitro assays. The KM values for oxygen have been determined via a range of in vitro experiments. The KM obtained is affected by substrate, conditions, and technique, and therefore provides only a possible indication of oxygen sensing capability. We have assessed, from the reported KMs and cellular data, whether these 2-OGDs might act as oxygen sensors. * Median KM from multiple studies (individual values listed in Supplementary Table S1).
2-OGD TypeEnzymeO2 KM (µM)Potential O2 Sensor (Yes/No)SubstrateEffect on Gene Expression in Hypoxia
HydroxylasesPHD1230 [17]YMultipleY
PHD2240 * [12,17,18,19]YMultiple
PHD3230 [17]YMultiple
4-PHα140 [17] Collagen
PAHX93 ± 43 [12,17]YIsovaleryl CoA
CDO176 ± 17 [12]NTaurine
FIH110 ± 30 * [12,20,21]Y/NMultipleY
Hydroxylases (known as DNA demethylases)TET130 ± 10 [22]NDNAY
TET230 ± 3 [22]NDNAY
JmjC demethylasesKDM4A60 ± 20 * [23,24,25]YH3K9me2/me3, H3K36me2, H1.4K26me2/me3Y
KDM4B150 ± 40 [25]YH3K9me2/me3, H3K36me2, H1.4K26me2/me3Y
KDM4C158 ± 13 [24]YH3K9me2/me3, H3K36me2, H1.4K26me2/me3Histone H3Y
KDM4E197 ± 16 [24]YH3K9me3Histone H3Y
KDM5A90 ± 30 [25]YH3K4me2/me3Histone H3Y
KDM5B40 ± 10 [25]NH3K4me2/me3Histone H3Y
KDM5C35 ± 10 [25]NH3K4me2/me3Histone H3Y
KDM5D25 ± 5 [25]NH3K4me2/me3Histone H3Y
KDM6A180 ± 40 [25]YH3K27me2/me3Histone H3Y
KDM6B20 ± 2 [25]NH3K27me2/me3Histone H3Y
Cysteamine (2-aminoethanethiol) dioxygenaseADO>500 [26]YRGS4 and RGS5
Table 2. Available mice and human mutations phenotypes for HIF and dioxygenases.
Table 2. Available mice and human mutations phenotypes for HIF and dioxygenases.
Gene
(Mouse/Human)
Homozygote Phenotype in MouseHuman Phenotype
HIFs
Hif1a/HIF1A (HIF-1α)Embryonic lethal with cardiovascular malformations, cephalic vascularisation and neural tube defects [27,28,29]Schizophrenia [30]. Maximal oxygen consumption [31]. Renal cell carcinoma [32].
Epas1/EPAS1 (HIF-2α)Embryonic lethal with bradycardia due to defective catecholamine homeostasis [33], vascular remodelling defects [34], cardiac failure and neonatal lethal with respiratory failure [35].Congenital heart disorder [36]. Autism spectrum disorder [37]. Pheochromocytoma/paraganglioma-polycythaemia [38,39,40,41,42,43,44] /somatostatinoma [45]. Erythrocytosis and polycythaemia with paraganglioma [40,41,42]. Erythrocytosis [46,47,48,49,50,51]. Pulmonary arterial hypertension [52].
Hif3a/HIF3A (HIF-3α)Mice deficient of an alternative spliced protein of HIF-3α, NEPAS, are viable and develop enlarged right ventricular owing to impaired pulmonary remodelling [53].NR
2-OGDs—hydroxylases
Egln2/EGLN2 (PHD1)Viable [54].Increased risk of hepatocellular carcinoma [55], lung cancer [56,57], gastric cancer [58], colorectal cancer [59]. Pheochromocytoma/paraganglioma-polycythemia [60].
Egln1/EGLN1 (PHD2)Embryonic lethal with severe cardiac and placental defects [54].High-altitude adaptation [45]. Erythrocytosis [61,62,63,64,65,66,67,68,69,70,71]. Pheochromocytoma/paraganglioma-polycythemia [60]. Pheochromocytoma [39]. Cardiopulmonary [72].
Egln3/EGLN3 (PHD3)Viable [54] with developmental defect of sympathoadrenal system [73].NR
P4ha1/P4HA1Embryonic lethal with delayed development and defective collagen IV assembly, resulting in base membrane rupture [74].Congenital-onset disorder of connective tissue [75].
P4ha2/P4HA2Viable and fertile with no obvious phenotypic abnormalities [76].High myopia [77].
Phyh/PHYH (PAHX)Viable without distinct developmental abnormalities [78].Refsum disease [79,80,81,82,83,84,85]. Nonsyndromic cleft lip and palate [86].
Hif1an/HIF1AN (FIH)Abnormal energy metabolism with reduced body weight, elevated metabolic rate and hyperventilation [87].Colorectal cancer [88].
2-OGDs—hydroxylases (mediators of DNA demethylation)
Tet1/TET1Knockout of TET1 via 5’ coding sequence results in partial embryonic lethality in mice [89,90,91], with surviving female mice displaying decreased fertility and reduced ovary size due to meiotic abnormality [89,90]. Whereas, mice with knockout via deletion of the catalytic domain of TET1 are viable and fertile [91,92,93], with slightly reduced body size [92], as well as impaired spatial learning and short-term memory [94].NR
Tet2/TET2Disordered hematopoiesis and eventually develop myeloid malignancies [95,96,97], and T- and B-cell malignancies [96].Myelodysplastic/myeloproliferative disease [98]. Prostate cancer [99]. Myeloproliferative neoplasms [100].
Tet3/TET3Neonatal lethality [90,101].Intellectual disability, developmental delay, autistic traits, hypotonia, growth abnormalities, facial dysmorphism and movement disorders [102].
2-OGDs—hydroxylases (RNA demethylases)
Fto/FTOAbnormal brain and cardiac development [103].Developmental delay and dysmorphic facial features [104]. Growth retardation and multiple malformations [103]. Developmental delay and growth retardation [105]. Growth retardation and multiple malformations [106]. Obesity [107,108]. Type II diabetes [109]. Metabolic syndrome including obesity, hypertension, dyslipidemia, and defective glucose tolerance [110].
2-ODGs—JmjC demethylases and hydroxylases
Jmjd4/JMJD4Viable and fertile with normal physiology [111].NR
Jmjd6/JMJD6Perinatal lethal with growth retardation and exhibit severe tissue and organ differentiation defects, including brain, lung, liver, kidney, intestine, heart and thymus development at different stages of embryogenesis [112,113,114,115].NR
Kdm2a/KDM2AEmbryonic lethal with severe growth retardation and defective neural tube closure [116].NR
Kdm2b/KDM2BKDM2B-1-deletion mice display moderate penetrance of neural tube defects, leading to exencephaly and death at birth [117]. Whereas, mice deficient of both KDM2B-1 and KDM2B-2 isoforms are embryonic lethal with fully penetrant developmental defects, including abnormal somitogenesis, reduced size, defective neural tube and heart [118,119,120]; especially a more severe developmental defect in female embryos [119]. Furthermore, KDM2B-2-deleted mice display similar developmental abnormalities with increased lethality, particularly in females [119].NR
Kdm3a/KDM3ADevelop obesity, abnormal fat metabolism [121,122], reduced energy expenditure, and display metabolic syndrome, including, high plasma cholesterol, insulin, triglyceride, and leptin levels [122]. Male infertility, smaller testes and severe oligozoospermia [123]. Retarded mammary gland ductal growth in female knockout mice [124].Male infertility [125].
Kdm3b/KDM3BPostnatal growth restriction and female mice were infertile due to decreased ovulation, prolonged estrous cycles, reduced fertilisation and uterine decidual response [126]. Male knockout mice have impaired reproductive function, sperm development and maturation [126]. Knockout mice exhibit myelodysplastic syndrome and defective hematopoiesis including leukocytosis, moderate anemia, and granulocytosis [127].Schizophrenia [128]. Intellectual disability [129]. Wilms tumour and hyperpigmentation [130]. Hepatoblastoma, autism, intellectual disability, and abnormal pigmentation [130]. Acute myeloid leukemia, mild intellectual disability, congenital hypothyroidism and congenital hip dysplasia [130]. Hodgkin lymphoma, feeding difficulties, intellectual disability, umbilical and inguinal hernia [131]. Intellectual disability, facial dysmorphism and short stature [131].
Jmjd1c/JMJD1CMales gradually develop infertility with decreasing testes size due to progressive loss of germ cells [127].Congenital heart disease in patients with 22q11.2 deletion syndrome [132]. Rett syndrome [133,134]. Autism spectrum disorder [133]. Intellectual disability [133]. Intracranial germ cell tumour [58].
Kdm4a/KDM4AViable [135].NR
Kdm4b/KDM4BViable [136]. Viable with lower birth rate. Early weaning results in death. Susceptible to obesity with impaired energy expenditure, adaptive thermogenesis and adipose tissue lipolysis [137].NR
Kdm4c/KDM4CViable and fertile [138]. However, another reported that it leads to embryonic lethality [139].Upper aerodigestive tract cancer [140]. Age at menarche [141].
Kdm4d/KDM4BViable and fertile without gross abnormalities [142].NR
Kdm5a/KDM5AViable [143,144]. Mice displayed mild behavioural and haematological abnormalities [143].Intellectual disability [145]. Congenital heart disease [146].
Kdm5b/KDM5BEmbryonic lethal [147,148]. Neonatal lethal due to failure to establish respiratory function, defective neural system and homeotic skeletal transformations [149].Intellectual disability, dyslexia, global developmental delay, facial dysmorphism, aggressive behaviour, hypospadias [150].
Kdm5c/KDM5CHemizygous KDM5C null male mice are embryonic lethal due to defective neurulation and cardiogenesis [151]. Male hemizygous knockout mice (Kdm5c−/y) Viable with adaptive and cognitive abnormalities, including increased aggression, impaired social behaviour, limited learning, fear memory deficits, defective dendritic spines [152,153] and significant reduced body weight [153].X-linked intellectual disability [154,155,156,157,158,159,160,161,162,163,164,165,166]. Autism spectrum disorder [167].
Kdm5d/KDM5DA large scale screening using CRISPR/Cas9-mediated genome editing reveals normal reproductive system in hemizygous KDM5D-knockout male mice [168].NR
Kdm6a/KDM6AEmbryonic lethal with cardiac development defects and neural tube closure. While female knockout mice died mid-gestational, some hemizygous KDM6A-null male mice survive into adulthood [151,169,170,171,172] and are fertile [171,172], with reduced lifespan and smaller size [171]. Female embryonic lethal, abnormal/truncated posterior bodies, anaemic (hematopoiesis), severe heart development defect and neural tube closure. Male died around birth due to neural tube closure defect and inability to breath [173]. Kabuki syndrome [174,175,176,177,178,179,180,181,182,183]. Biliary atresia with Kabuki syndrome-like features [184]. Renal cancer [185].
Kdm6b/KDM6BEmbryonic [186] and perinatal lethal with respiratory failure [187,188,189], detail reviewed here [190]. Reduced proliferation and hypertrophy of chondrocytes, as well as delayed endochondral ossification in mice [191]. Delayed osteoblast differentiation and bone ossification [192].Intellectual disability [145]. Intellectual disability, brachydactyly and dysmorphism [193].
Uty/UTYHemizygous male mice are viable [169].NR
Kdm7a/KDM7AA large-scale genome-wide tissue phenotype screen revealed that abnormal hair follicles, sebaceous gland, tail and hair follicle bulge morphology in KDM7A knockout mice [194].
Phf8/PHF8Impaired learning and memory, hippocampal long-term potentiation [195].X-linked mental retardation with cleft lip/palate [196,197,198]. Autism and Asperger syndrome [199]. Autism spectrum disorder, intellectual disability, cleft palate and Aarskog syndrome [200]. Intellectual disability [159].
Kdm8/KDM8Embryonic lethal with delayed development in multiple organs [201] and growth retardation [202].NR
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Frost, J.; Frost, M.; Batie, M.; Jiang, H.; Rocha, S. Roles of HIF and 2-Oxoglutarate-Dependent Dioxygenases in Controlling Gene Expression in Hypoxia. Cancers 2021, 13, 350. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers13020350

AMA Style

Frost J, Frost M, Batie M, Jiang H, Rocha S. Roles of HIF and 2-Oxoglutarate-Dependent Dioxygenases in Controlling Gene Expression in Hypoxia. Cancers. 2021; 13(2):350. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers13020350

Chicago/Turabian Style

Frost, Julianty, Mark Frost, Michael Batie, Hao Jiang, and Sonia Rocha. 2021. "Roles of HIF and 2-Oxoglutarate-Dependent Dioxygenases in Controlling Gene Expression in Hypoxia" Cancers 13, no. 2: 350. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers13020350

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