Responding appropriately to changes in oxygen availability is essential for multicellular organism survival. Molecularly, cells have evolved intricate gene expression programmes to handle this stressful condition. Although it is appreciated that gene expression is co-ordinated by changes in transcription and translation in hypoxia, much less is known about how chromatin changes allow for transcription to take place. The missing link between co-ordinating chromatin structure and the hypoxia-induced transcriptional programme could be in the form of a class of dioxygenases called JmjC (Jumonji C) enzymes, the majority of which are histone demethylases. In the present review, we will focus on the function of JmjC histone demethylases, and how these could act as oxygen sensors for chromatin in hypoxia. The current knowledge concerning the role of JmjC histone demethylases in the process of organism development and human disease will also be reviewed.

Changes to oxygen availability or increased oxygen demand create an imbalance called hypoxia. Hypoxia is an important physiological stimulus for embryo development of mammals, but it is also a serious component of the pathology of many human diseases [13]. Given its importance, it has attracted a great amount of research, which has helped to understand how the physiology of oxygen sensing and response works. However, at the cellular and molecular level, great unknowns still exist. The molecular mechanisms underlying the cellular response were significantly boosted by the discovery of the main transcription factor family controlled by oxygen, called HIF (hypoxia-inducible factor) in the early 1990s [4]. HIF is now known to be a heterodimer of the oxygen-controlled subunit HIF-α and the oxygen-insensitive subunit HIF-1β, which was originally identified as a binding partner for the aryl hydrocarbon receptor, and as such has the gene name of ARNT (aryl hydrocarbon nuclear translocator) [5].

In mammalians there are three genes for HIF-α subunits, HIF-1α, HIF-2α (gene name EPAS1, for endothelial Pas protein 1) and HIF-3α [6]. HIF-α is controlled by oxygen post-translationally via the action of dioxygenase enzymes such as prolyl hydroxylases and FIH (factor inhibiting HIF). Prolyl hydroxylase-mediated hydroxylation of HIF α-subunits creates a high-affinity binding site for the ubiquitin ligase complex containing the tumour suppressor VHL (von Hippel–Lindau protein) [710]. VHL is part of the cullin-2, elongin B/C and the small RING finger protein RBX1 ligase complex [11], which promotes Lys48-linked ubiquitination, and hence proteasomal-mediated degradation. FIH-mediated hydroxylation produces an inhibitory moiety in the C-terminal transactivation domain of HIF-1α and HIF-2α, preventing the binding of co-activator proteins such as p300 and hence reducing the transcriptional activity of the transcription factor [12].

HIF-dependent genes are varied and many, with over 100 direct genes being validated and several more putative genes recently identified by genomic approaches, such as ChIP-sequencing [13]. Genes involved in restoration of oxygen homoeostasis, cell survival and growth as well as metabolism have received great interest from the medical community, as these can be used not only as biomarkers, but also as direct therapeutic targets for diseases such as cancer [14].

One interesting new class of HIF-dependent targets are the dioxygenase enzymes called JmjC (Jumonji C)-containing proteins [15]. These enzymes are, for the most part, protein demethylases, and were identified as being structurally similar to the HIF hydroxylase FIH [16,17]. On the basis of the structural analysis of these enzymes, it became clear that these enzymes, much like FIH, would require molecular oxygen and 2-oxoglutarate to catalyse their enzymatic reactions, suggesting that JmjC enzymes could act as molecular oxygen sensors in the cell. In the present review, we will discuss the evidence available on how JmjC histone demethylases can contribute to oxygen sensing by altering chromatin structure and function.

Chromatin is the collective name for DNA and the protein complexes, of which the nucleosome (DNA and histone octamer) is the basic unit [18]. Initially thought as a passive impediment to nuclear processes, it is now known that chromatin is highly dynamic, and responsive to several stimuli and stages of development [19]. Two major states of chromatin are generally accepted: heterochromatin (compact/silent) and euchromatin (open/active) [18]. However, imaging approaches have revealed additional distinct compaction levels do exist in both types of chromatin domains [20].

Chromatin structure can be altered by several mechanisms, either involving ATP-dependent remodellers [21] or changes to the histone octamer [22,23], including alternative histone and histone post-translation modifications. The number of possible histone modifications is vast; however, one such modification, with relevance to the present review, is histone methylation. Histone methylation can occur on several lysine or arginine residues, primarily on histone H3 and histone H4 [24]. Lysine methylation, unlike acetylation, does not change the charge of the protein. As such, methylation marks change chromatin structure by different mechanisms, involving the recruitment or inhibition of recruitment of distinct enzymatic complexes [25,26]. Functionally, histone methylation can both activate and repress transcription [25] as well as control DNA replication [26]. Histone methylation can lead to compaction or relaxation of chromatin (Figure 1), depending on which histone residue is methylated and, as such, how the recruitment of the corresponding enzymatic complex is altered [22].

Histone methylation marks associate with chromatin compaction status

Figure 1
Histone methylation marks associate with chromatin compaction status

Active chromatin or euchromatin is associated with a more open status and is characterized by several types of histone methylation marks as depicted. On the other hand, silent chromatin or heterochromatin is associated with closed and compacted status, also characterized by a subset of histone methylation patterns as depicted. JmjC enzymes can thus, in principle, control chromatin status by removing the histone methyl groups.

Figure 1
Histone methylation marks associate with chromatin compaction status

Active chromatin or euchromatin is associated with a more open status and is characterized by several types of histone methylation marks as depicted. On the other hand, silent chromatin or heterochromatin is associated with closed and compacted status, also characterized by a subset of histone methylation patterns as depicted. JmjC enzymes can thus, in principle, control chromatin status by removing the histone methyl groups.

Close modal

Soon after its identification, histone methylation was thought to be a stable modification, as no enzymes were known to remove this mark. However, now it is known that histone methylation can be removed by two different classes of enzymes: the LSD (lysine-specific demethylase) family and the JmjC family (Table 1). For a more detailed review of the LSD family, please see [27].

Table 1
JMJC-containing protein in humans according to UniProt

Y, yes; N, no.

Gene nameUniProt numberAdditional namesHistone demethylase activity (Y/N)
HR O43593 Hairless, HAIR 
HSPBAP1 Q96EW2 HBAP1 
HIF1AN Q9NWT6 HIF1N, FIH 
JARID2 Q92833 JARD2 N* 
JMJD1C Q15652 JHD2C N 
JMJD4 Q9H9V9 
JMJD6 Q6NYC1 PSR Y 
JMJD7 P0C870 
JMJD8 Q96S16 
KDM2A Q9Y2K7 FBXL11, JHDM1A 
KDM2B Q8NHM5 FBXL10, JHDM1B 
KDM3A Q9Y4C1 JMJD1A 
KDM3B Q7LBC6 JMJD1B 
KDM4A O75164 JMJD2A 
KDM4B O94953 JMJD2B 
KDM4C Q9H3R0 JMJD2C 
KDM4D Q6B0I6 JMJD2D 
KDM4E B2RXH2 JMJD2E 
KDM5A P29375 RBBP2, JARID1A 
KDM5B Q9UGL1 PLU1, JARID1B 
KDM5C P41229 JARID1C 
KDM5D Q9BY66 JARID1D 
KDM6A O15550 UTX 
KDM6B O15054 JMJD3 
JHDM1D Q6ZMT4 KDM7A 
KDM8 Q8N371 JMJD5 
MINA Q8IUF8 MINA53, NO52 Y§ 
NO66 Q9H6W3 C14orf169 
PHF2 O75151 CENP35, JHDM1E 
PHF8 Q9UPP1 JHDM1F 
TYW5 A2RUC4 
UTY O14607 
Gene nameUniProt numberAdditional namesHistone demethylase activity (Y/N)
HR O43593 Hairless, HAIR 
HSPBAP1 Q96EW2 HBAP1 
HIF1AN Q9NWT6 HIF1N, FIH 
JARID2 Q92833 JARD2 N* 
JMJD1C Q15652 JHD2C N 
JMJD4 Q9H9V9 
JMJD6 Q6NYC1 PSR Y 
JMJD7 P0C870 
JMJD8 Q96S16 
KDM2A Q9Y2K7 FBXL11, JHDM1A 
KDM2B Q8NHM5 FBXL10, JHDM1B 
KDM3A Q9Y4C1 JMJD1A 
KDM3B Q7LBC6 JMJD1B 
KDM4A O75164 JMJD2A 
KDM4B O94953 JMJD2B 
KDM4C Q9H3R0 JMJD2C 
KDM4D Q6B0I6 JMJD2D 
KDM4E B2RXH2 JMJD2E 
KDM5A P29375 RBBP2, JARID1A 
KDM5B Q9UGL1 PLU1, JARID1B 
KDM5C P41229 JARID1C 
KDM5D Q9BY66 JARID1D 
KDM6A O15550 UTX 
KDM6B O15054 JMJD3 
JHDM1D Q6ZMT4 KDM7A 
KDM8 Q8N371 JMJD5 
MINA Q8IUF8 MINA53, NO52 Y§ 
NO66 Q9H6W3 C14orf169 
PHF2 O75151 CENP35, JHDM1E 
PHF8 Q9UPP1 JHDM1F 
TYW5 A2RUC4 
UTY O14607 
*

No histone demethylase activity, but modulates methyltransferases.

No evidence found yet.

Only in vitro.

§

Not the main enzymatic activity.

On the basis of the presence of the JmjC domain, there are currently 32 proteins in humans belonging to this class (Table 1). However, not all of these have been associated with histone demethylation. Although some of the JmjC proteins without histone demethylase activity have important biological functions in the cell, the present review will focus on the regulation of JmjC histone demethylases. As mentioned above, the JmjC class of proteins require oxygen and 2-oxoglutarate for their activity. Elegant structural work from the Schofield group and other laboratories has greatly helped in the understanding of how target selectivity is achieved, as well as with the general mechanism of demethylation by these enzymes [24,2830].

For catalytic activity, JmjC histone demethylases use molecular oxygen and 2-oxoglutarate with release of CO2, succinate and formaldehyde [29,31]. As such, reduction in the availability of oxygen would have an impact on catalytic activity. Work performed using recombinant prolyl hydroxylases and FIH proteins has shown that these enzymes have different Km values for oxygen [32]. Although the prolyl hydroxylase Km value is approximately 230 μM O2, the FIH Km value for O2 is only 90 μM [32]. These results indicated that prolyl hydroxylases are inhibited with a smaller reduction in the availability of O2 than FIH. However, thus far oxygen affinity for the majority of the JmjC enzymes has not been determined. One enzyme that has been investigated, which belongs to the KDM4 (lysine-specific demethylase 4) family, is KDM4E [33]. It was shown that KDM4E reacts slowly with O2, at a similar level to prolyl hydroxylase 2. It was also suggested that KDM4E has an incremental response over physiologically relevant ranges of O2 [33]. This analysis did indicate the potential of these enzymes to act as oxygen sensors in the cell. However, it would be necessary for more biochemical studies to be performed in vitro to really establish, and compare, the oxygen requirements of JmjC histone demethylases with other known dioxygenases such as prolyl hydroxylases and FIH.

An additional regulation of JmjC demethylases by oxygen, albeit indirect, is via transcriptional regulation. Transcriptional analyses, using several different cellular systems, have shown that a great number of JmjC histone demethylases are hypoxia- inducible at the mRNA levels [34]. These include KDM2A, KDM2B, KMD3A, KDM3B, KMD4B, KDM4C, KDM4D, KDM5A, KDM5A, KDM5B, KDM5C, KDM5D, KDM6A, KDM6B, KDM8, JARID2 and PHF8 (plant homeodomain finger protein 8) (references in [34]). Some of these enzymes have been shown to be direct targets of HIF-1α. The HIF-1α-regulated JmjC histone demethylases are: KDM3A [3538], KDM4B [35,36], KDM4C [36], KDM5C [39] and KDM6B [40]. Whether any of the additional JmjC histone demethylases that were found to be hypoxia-inducible are also HIF-dependent remains to be investigated. In addition, their regulation by HIF-2 has not been directly investigated, apart from KDM3A, KDM4B and KDM4C, which are mainly regulated by HIF-1α [36]. As such, it is not known whether the remaining hypoxia-inducible JmjC enzymes are HIF-1-specific targets. Current studies should help elucidate these questions.

Histone methylation is probably one of the most studied chromatin modifications with a great number of studies describing the role of a specific methylation mark in the control of gene expression. The recent availability of large-scale and genomic sequencing data has also helped to associate different methylation marks with the diverse chromatin states [41]. Generally, methylation at Lys4, Lys36 or Lys79 of histone H3 are hallmarks of actively transcribed genes, whereas methylation of Lys9 and Lys27 of histone H3, as well as of histone H4 Lys20, are associated with transcriptional repression and heterochromatin formation [42] (Table 2). Mutual exclusiveness of these marks establishes the importance of histone demethylases in the remodelling of chromatin and reprogramming of gene expression.

Table 2
Methylation marks as a determinant of chromatin states
Methylation markAssociated activity
H3K4me2/3 Hallmark of regulatory elements at the 5′ end of transcriptionally active genes or of genes poised for transcriptional activation 
H3K4me1 Hallmark of enhancer sequences 
H3K36me1/2 Restricted to the body and 3′ end of the gene 
H3K36me2/3 and H3K79me2/3 H3K36me2/3 and H3K79me2/3 are enriched in gene bodies 
H3K9me3 and H4K20me3 Associate with non-genic regions, repetitive or transposable DNA elements including satellite sequences and long terminal repeats 
H3K4me3 and H3K27me3 Bivalent domain in embryonic stem cells associated with complex transcription 
Methylation markAssociated activity
H3K4me2/3 Hallmark of regulatory elements at the 5′ end of transcriptionally active genes or of genes poised for transcriptional activation 
H3K4me1 Hallmark of enhancer sequences 
H3K36me1/2 Restricted to the body and 3′ end of the gene 
H3K36me2/3 and H3K79me2/3 H3K36me2/3 and H3K79me2/3 are enriched in gene bodies 
H3K9me3 and H4K20me3 Associate with non-genic regions, repetitive or transposable DNA elements including satellite sequences and long terminal repeats 
H3K4me3 and H3K27me3 Bivalent domain in embryonic stem cells associated with complex transcription 

Considering the role of the methylation marks and demethylase specificity, KDM2 and KDM5 families can promote the formation of repressed chromatin, KDM3, KDM6 and KDM7 act as chromatin activators and the KDM4 family may have different effects on chromatin status [41]. However, less is known about the interplay between histone demethylases and other important types of chromatin-remodelling enzymes, such as the ATP-dependent chromatin remodellers [CRCs (chromatin-remodelling complexes)].

Four subfamilies of CRCs have been characterized in mammals: SWI/SNF, CHD (chromodomain helicase DNA binding), Ino80 and ISWI (imitation-SWI protein) [34]. They share a similar ATPase domain responsible for the disruption of protein–DNA interactions in nucleosome using the energy of ATP hydrolysis. However, all four subfamilies specialize in particular purposes and biological context, depending on their interaction network [18,21]. The broad interaction network of chromatin remodellers is regulated by unique domains flanking the ATPase domain or by the presence of accessory subunits (Figure 2).

Schematic representation of chromatin-remodelling complexes, highlighting the domains present in different subunits

Figure 2
Schematic representation of chromatin-remodelling complexes, highlighting the domains present in different subunits

ARID, AT-rich interactive domain; HAND-SANT-SLIDE, DNA-binding domains; HSA, domain binding nuclear actin-related proteins (ARPs) and actin; NuRF, ISWI-related chromatin remodeller.

Figure 2
Schematic representation of chromatin-remodelling complexes, highlighting the domains present in different subunits

ARID, AT-rich interactive domain; HAND-SANT-SLIDE, DNA-binding domains; HSA, domain binding nuclear actin-related proteins (ARPs) and actin; NuRF, ISWI-related chromatin remodeller.

Close modal

There is very little information concerning how JmjC enzymes interact or control the action of CRCs. However, a few studies have suggested that demethylases can alter the action of such remodellers. LSD1 and several JmjC enzymes were found in a genetic screen in Drosophila as corepressors of SWI/SNF activity during wing development [43]. In addition, KDM6A was shown to interact with BRM, one of the catalytic helicases in the SWI/SNF complex, modulating the acetylation of H3K27 (where K indicates the lysine residue under investigation, i.e. H3K27 is Lys27 of histone H3) via binding with CBP (cAMP-response-element-binding protein-binding protein) [44].

Despite the lack of direct research investigating how JmjC control the action of CRCs, indirect evidence does exist to support this hypothesis. As such, ISWI- and CHD- remodelling complexes contain tandem CDs (chromodomains) and PHDs (plant homeodomains) that are able to recognize methylated lysine residues (Figure 3A). Histone methylation was shown to be important for the recruitment and stabilization of CRCs [4547]. Different methylation marks are associated with ISWI and CHD binding. They are recruited to promoter and enhancer elements enriched in H3K4 methylation, known to regulate transcription initiation [4547], recruited to methylated H3K36 in gene bodies linking remodellers to transcription elongation and termination [48] and recruited to the H3K9 methylated regions of repressed chromatin (Figure 3A). As such, given the importance of histone methylation for CRC recruitment, it would be rational to hypothesize that histone demethylases can regulate this process. Future research directed at this particular question would be necessary to fully establish how JmjC enzymes co-ordinate the action of CRCs.

How histone methylation controls CRC recruitment

Figure 3
How histone methylation controls CRC recruitment

(A) Histone methylation marks localization and chromatin remodellers associated with them. bromo, bromodomain; me, methylation site. (B) Regulation of NuRD recruitment by histone methylation marks. Set9, H3K9 methyltransferase.

Figure 3
How histone methylation controls CRC recruitment

(A) Histone methylation marks localization and chromatin remodellers associated with them. bromo, bromodomain; me, methylation site. (B) Regulation of NuRD recruitment by histone methylation marks. Set9, H3K9 methyltransferase.

Close modal

An interesting example of the action of CRCs controlled by histone methylation marks is the recognition of methylated lysine residues by the NuRD (nucleosome-remodelling deacetylase) complex (Figure 3B). NuRD is a repression complex that combines deacetylase activity with ATP-remodelling activity to form a compacted chromatin state, and thus repress gene expression. It contains a CHD3/CHD4 ATPase domain with tandem PHDs and CDs in its N-terminus. The PHD fingers were shown to interact with H3K9me3 [where me indicates the methylation status, from me1 (monomethylated) to me3 (trimethylated)] and this interaction can be regulated by the methylation status of H3K4 [49]. Methylation of H3K4 by Set9 reduces the association of the complex with the histone H3 tail, which is the most probable mechanism for regulating the formation of repressed chromatin by the NuRD complex [50]. The histone demethylase LSD1 has been linked with the NuRD complex [51], indicating that indeed histone demethylases can modulate CRC's activity. However, the involvement of JmjC in the control of NuRD recruitment has not been investigated thus far.

Almost all CRCs apart from CHD have been implicated in the hypoxia response [5255]. As such, and considering the level of cross-talk present in chromatin remodelling, more links between lysine demethylation and ATP-dependent chromatin remodelling will be discovered in the near future.

KDM2A was the first published JmjC-domain-containing protein shown to have histone demethylase activity [31]. Since then, many more JmjC histone demethylases have been uncovered and their histone targets investigated. In vitro studies have revealed that while some JmjC have quite particular selectivity for histone residues, others have a broader range of targets (Figure 4). However, which JmjC histone demethylase is active in cells is still unknown, and most likely will vary from cell type to cell type, as well as developmental state.

JmjC family domain structure and histone targets

Figure 4
JmjC family domain structure and histone targets

ARID, AT-rich interactive domain; C5HC2-ZF, C5HC2 zinc finger domain; CXXC-ZF, CXXC zinc finger domain; FBox, F-box domain; JmjC, Jumonji C domain; JmjN, Jumonji N domain; LRR, leucine-rich repeat domain; TPR, tetratricopeptide domain; Tudor, Tudor domain.

Figure 4
JmjC family domain structure and histone targets

ARID, AT-rich interactive domain; C5HC2-ZF, C5HC2 zinc finger domain; CXXC-ZF, CXXC zinc finger domain; FBox, F-box domain; JmjC, Jumonji C domain; JmjN, Jumonji N domain; LRR, leucine-rich repeat domain; TPR, tetratricopeptide domain; Tudor, Tudor domain.

Close modal

Studies in hypoxia have shown that histone methylation marks are indeed increased (Figure 5) [5658]. Exposure of Hepa1-6 cells to 0.2% O2 for 48 h induces global increases in H3K4me2, H3K4me3, H3K79me3, H3K27me3 and H3K9me2 [56]. Another study found that changes in histone methylation marks in mouse macrophage cells were only visible after 24 h of exposure when oxygen levels were below 3%. Furthermore, at 1% oxygen there was a global increase in H3K9me2, H3K9me3 and H3K36me3, and these changes were attributed to inhibition of histone demethylase activity [58]. These studies suggest that a variety of JmjC enzymes can be inhibited by low oxygen conditions; however, given that most of these enzymes can target several histone residues (Figure 4), additional work is required to really establish which enzymes are altered in hypoxia.

Hypoxia induces the increase of certain histone methylation marks

Figure 5
Hypoxia induces the increase of certain histone methylation marks

Diagram depicting the relationship between oxygen concentration and increased levels of histone methylation marks observed in several studies.

Figure 5
Hypoxia induces the increase of certain histone methylation marks

Diagram depicting the relationship between oxygen concentration and increased levels of histone methylation marks observed in several studies.

Close modal

One study has reported a global increase in H3K4me3 at 1% oxygen, as a result of KDM5A inhibition [57]. That study also observed an increase in H3K4me3 at HMOX1 (haem oxygenase 1) and DAF gene promoters [57], thus suggesting that global and local increases in H3K4me3 due to KDM5A inhibition during hypoxia may lead to altered gene transcription.

However, there is still very little information about the role of histone methylation on chromatin structure and JmjC-specific actions during hypoxia. One important question to answer is how are particular histone modifications associated with active and repressive transcription altered following hypoxic stress. In addition, dynamic analysis is also lacking, as the oxygen sensing and response system in cells is usually programmed to reset after prolonged hypoxia, as is observed in the regulation of HIF levels by prolyl hydroxylases. As such, it would be hypothesized that changes in histone methylation marks would be more dramatic and directly dependent on JmjC inhibition at earlier times of hypoxia exposure. Prolonged hypoxia would involve not only inhibition of these enzymes, but also, and most likely, compensatory mechanisms by which increased expression of JmjC proteins would occur.

Although cell culture studies have revealed important information on the complex molecular functions, regulation and targets of JmjC histone demethylases, whole organism studies are vital in understanding the biological significance of these enzymes in vivo. The importance of JmjC demethylases in biological processes is illustrated by their association with diseases and depletion studies in model organisms (Table 3), showing that they have key roles in development. Studies in invertebrate model organisms found developmental defects when certain JmjC enzymes are depleted. For example, knockout of two PHF family orthologues in zebrafish, Jhdm1da and Jhdm1db, leads to abnormal posterior development [59]. Also knockdown of Lid, the KDM5 family orthologue in Drosophila, causes numerous developmental defects through deregulation of homeotic gene expression [60].

Table 3
Phenotypes associated with JmjC depletion studies in model organisms

–, no orthologue; ND, not determined or no data.

Phenotype in depletion studies
JmjCMus musculusDanio rerioCaenorhabditis elegansDrosophila melanogaster
KDM2B Partial peri/postnatal lethal, lowered sperm count, defective neural tube development [112ND ND Embryonic lethal [113
KDM3A Metabolic defects, adult obesity and male infertility [114116– – ND 
KDM4A Impaired cardiac stress response [60ND Increased germline apoptosis [117Abnormal male wing extension and reduced male lifespan [118
KDM4D No detected phonotypical change [119ND   
KDM5A Decreased apoptosis in HSC and myeloid progenitors [120ND Defects in vulva formation and reduced lifespan [83,121Developmental defects [122
KDM5B Embryonic lethal [123ND   
KDM5C Cardiac looping and neuralation defects [124Increased neuronal cell death and abnormal dendrite development [90  
JARID2 Embryonic lethal [125127ND – ND 
KDM6B Perinatal lethal and defective lung development [128ND Defective vulva development [129Larval lethal [130
KDM6A Partial male embryonic lethal, defects in neural tube and cardiac development and female embryonic lethal [63,124Abnormal posterior development [59  
PHF2 Partial neonatal lethal, adipogenesis defects [61Detects in brain development [131Defects in body movement [97– 
KDM8 Embryonic lethal [132ND ND ND 
JMJD6 Perinatal lethal with multiple developmental defects [133Defects in heart brain, somites and notochord [134Mild cell engulfment defects [135Enhanced apoptosis in developing eye [136
Phenotype in depletion studies
JmjCMus musculusDanio rerioCaenorhabditis elegansDrosophila melanogaster
KDM2B Partial peri/postnatal lethal, lowered sperm count, defective neural tube development [112ND ND Embryonic lethal [113
KDM3A Metabolic defects, adult obesity and male infertility [114116– – ND 
KDM4A Impaired cardiac stress response [60ND Increased germline apoptosis [117Abnormal male wing extension and reduced male lifespan [118
KDM4D No detected phonotypical change [119ND   
KDM5A Decreased apoptosis in HSC and myeloid progenitors [120ND Defects in vulva formation and reduced lifespan [83,121Developmental defects [122
KDM5B Embryonic lethal [123ND   
KDM5C Cardiac looping and neuralation defects [124Increased neuronal cell death and abnormal dendrite development [90  
JARID2 Embryonic lethal [125127ND – ND 
KDM6B Perinatal lethal and defective lung development [128ND Defective vulva development [129Larval lethal [130
KDM6A Partial male embryonic lethal, defects in neural tube and cardiac development and female embryonic lethal [63,124Abnormal posterior development [59  
PHF2 Partial neonatal lethal, adipogenesis defects [61Detects in brain development [131Defects in body movement [97– 
KDM8 Embryonic lethal [132ND ND ND 
JMJD6 Perinatal lethal with multiple developmental defects [133Defects in heart brain, somites and notochord [134Mild cell engulfment defects [135Enhanced apoptosis in developing eye [136

Several JmjC histone demethylases have been depleted in mice (Table 3). Phenotypes of these knockouts have large variation, ranging from embryonic lethality in KDM8, JARID2 and KDM5B knockouts to male infertility and obesity in the KDM3A knockout, to no detectable phenotypical change in KDM4D depletion. This demonstrates that, although some of these demethylases are essential in development, functional redundancy partly as a result of overlapping targets is likely to account for milder phenotypes observed from depletion of others. Mechanistic insights have been gained from some of these studies. For example, PHF2-knockout mice have reduced adipose tissue, PHF2 interacts with the key regulator of adipogenesis CEBPA and is proposed to promote adipogenesis through co-activation of CEBPA [61,62]. This has led to the suggestion that PHF2 may be a novel therapeutic target for the treatment of obesity and metabolic diseases.

The cardiac defects observed in KDM6A-depleted mice are likely to be due to deregulation of cardiac gene expression programmes, as KDM6A is recruited to cardiac-specific enhancer regions where it associates with multiple transcription factors [63]. However, for several of the depletion studies performed, molecular mechanisms to explain the observed phenotypes are still ill-defined and will require further investigation. Furthermore, for many of the JmjCs, there is limited information from in vitro studies and no/limited information from in vivo studies, which will need to be addressed in order to characterize these enzymes.

One aspect that highlights the importance of a particular gene is its association with human disease. As such, there are numerous links between JmjC histone demethylases and human diseases, most notably cancer and neurological disorders (Figure 6). Genetic alterations in JmjCs, including KDM2A, KDM2B, KDM4A, KDM4B, KDM42C and KDM5B, have all been linked to cancer progression and these enzymes have been suggested as chemotherapeutic targets [6473].

JmjC alterations in human disease

Figure 6
JmjC alterations in human disease

Diagram of the observed associations of JmjC expression in different human diseases, highlighting KDM4 as particularly susceptible to deregulation in a variety of human pathologies.

Figure 6
JmjC alterations in human disease

Diagram of the observed associations of JmjC expression in different human diseases, highlighting KDM4 as particularly susceptible to deregulation in a variety of human pathologies.

Close modal

The KDM2 group of JmjCs appear to have both cancer-promoting and -inhibiting functions depending on cellular context, through their regulation of cell proliferation. In cell culture, NF-κB (nuclear factor κB)-dependent colon cancer cell growth is impaired by KDM2A [74] and KDM2B has been demonstrated to negatively regulate cell proliferation via repression of ribosomal RNA genes [75]. Conversely, overexpression of KDM2A and KDM2B in mouse embryo fibroblasts confers resistance to stress-induced senescence [66]. KDM2B positively regulates cell proliferation and growth through silencing of the cell cycle inhibitor p15Ink4b [66,76]. KDM2A is overexpressed in a subset of non-small cell lung cancer patients and has been reported to drive cancer progression in these tumours through up-regulation of ERK1/2 (extracellular-signal-regulated kinase 1/2) signalling [64]. KDM2B has been identified both as oncogene and tumour suppressor through proviral insertional mutagenesis studies in rodents [77,78]. Various leukaemias and bladder cancers display up-regulated KDM2B [65,66] and it is proposed to play a key role in leukaemia progression through loss of p15Ink4b function [65]. As well as being overexpressed in certain cancers, KDM2A and KDM2B have also been found to be down-regulated in prostate cancer and glioblastoma respectively [75,79]. These studies have highlighted the cell-dependent nature of the KDM2 family's mode of action.

The KDM4 group has also been associated with various cancers. KDM4A is down-regulated in bladder cancer [80], and KDM4A and KDM4B are up-regulated in breast cancer and peripheral nerve sheath tumours respectively [67,81]. KDM4B is oestrogen inducible and has been shown to promote oestrogen-stimulated breast cancer proliferation [82]. Amplification of KDM4C has been reported in breast cancer, oesophageal squamous cell cancer and medulloblastoma [68,69,83] and a translocation involving KDM4C is present in mucosa-associated lymphoid tissue lymphomas [70]. KDM4C overexpression induces transformation in immortalized mammalian epithelial cell lines, suggesting a role for KDM4C in driving tumorigenesis [83]. This is supported by studies showing that knockdown of KDM4C limits proliferation in mammalian cancer cell lines [8385].

The KDM5 group also has strong association with human diseases. KMD5A is associated with tumourigenesis in the lung [86] and is also associated with haemopoietic malignancies [87]. KDM5B has low expression levels in normal adult tissue, except in the testes, but it is found overexpressed in the bladder, prostate and breast cancer [7173]. KDM5B is a transcriptional co-activator of the androgen receptor, but can also function as a transcriptional repressor [88]. It has also been shown to promote breast cancer proliferation in both in vitro and in vivo. Depletion of KDM5B limits growth in MCF-7 breast cancer cells and in a mouse breast cancer model, and this correlates with repression of tumour suppressor genes including BRCA1 (breast cancer early-onset 1) [89].

Several JmjCs, including KDM5C and PHF8, have strong links to neurological development and defects. KDM5C is important for neuronal survival in primary mammalian neurons and dendritic development in zebrafish [90]. Furthermore, mutations in KDM5C are frequently found in X-linked mental retardation [9194]. KDM5C is involved in REST (repressor element 1-silencing transcription factor)-mediated transcriptional repression and loss of KDM5C activity, leading to deregulation of neuronal genes under transcriptional regulation of REST, which has been proposed as the mechanism between KDM5C and X-linked mental retardation [95]. PHF8 loss-of-function mutations have also been found in X-linked mental retardation patients as well as patients with cleft palate [96], it is suggested that KDM5C and other X-linked mental retardation-associated genes are under regulation by PHF8 via the transcription factor ZNF711 (zinc finger protein 711) [97].

KDM6A has been associated both with cancer and X-linked syndromes such as Kabuki syndrome [98101]. Similarly, KMD6B contributes to tumour suppression by activation of p14ARF [102], as well as co-operating with p53 [103]. It has also been associated with pro-inflammatory gene expression in immune cells [104106].

Epigenetic deregulation is a key driver of tumorigenesis and neurological disorders as well as other diseases. The pathological roles of enzymes modifying histone methylation, including JmjC histone demethylases, is only beginning to be understood and these enzymes provide promising drug targets. As such, several studies have already attempted to employ structural and medicinal chemical strategies to target these enzymes for therapy [104,107111]. As JmjC histone demethylases require oxygen as a cofactor, it is possible to speculate that histone marks can be rapidly altered when hypoxia is present. This would indicate that chromatin structure would change and adapt to low oxygen, possibly even more rapidly than any other process in the cell, and would thus set the landscape for the hypoxia-induced transcriptional program observed in many cells. However, additional work regarding the specific requirements for oxygen for each of the JmjCs is required to fully verify this hypothesis. As more information is gathered regarding the molecular function of the individual JmjC enzymes, more targeted approaches can be employed. Thus there is still great potential for using JmjC enzymes as valid new targets for therapy in human disease. Future research analysing the role of individual JmjC enzymes, and their relationship with chromatin-remodelling complexes, should provide exciting and useful insights into the biology of JmjC enzymes as well as exploit their potential for therapeutic targeting.

CD

chromodomain

CHD

chromodomain helicase DNA binding

CRC

chromatin-remodelling complex

FIH

factor inhibiting HIF

HIF

hypoxia-inducible factor

ISWI

imitation-SWI protein

JmjC

Jumonji C

KDM

lysine-specific demethylase

LSD

lysine-specific demethylase

NuRD

nucleosome-remodelling deacetylase

PHD

plant homeodomain

PHF

PHD finger protein

REST

repressor element 1-silencing transcription factor

VHL

von Hippel–Lindau protein

This study was funded by a Wellcome Trust Doctoral training fellowship (to A.S.), a Medical Research Council doctoral training studentship (to M.B.) and a Cancer Research UK Senior Fellowship [grant number C99667/A12918 (to S.R.)], with additional funding from the Wellcome Trust [grant number 097945/B/11/Z] and Tenovus Scotland.

1
Mole
 
D. R.
Ratcliffe
 
P. J.
 
Cellular oxygen sensing in health and disease
Pediatr. Nephrol.
2008
, vol. 
23
 (pg. 
681
-
694
)
[PubMed]
2
Ratcliffe
 
P. J.
 
Oxygen sensing and hypoxia signalling pathways in animals: the implications of physiology for cancer
J. Physiol.
2013
, vol. 
591
 (pg. 
2027
-
2042
)
[PubMed]
3
Semenza
 
G. L.
 
Oxygen sensing, hypoxia-inducible factors, and disease pathophysiology
Annu. Rev. Pathol.
2014
, vol. 
9
 (pg. 
47
-
71
)
[PubMed]
4
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
, vol. 
12
 (pg. 
5447
-
5454
)
[PubMed]
5
Bersten
 
D. C.
Sullivan
 
A. E.
Peet
 
D. J.
Whitelaw
 
M. L.
 
bHLH-PAS proteins in cancer
Nat. Rev. Cancer
2013
, vol. 
13
 (pg. 
827
-
841
)
[PubMed]
6
Rocha
 
S.
 
Gene regulation under low oxygen: holding your breath for transcription
Trends Biochem. Sci.
2007
, vol. 
32
 (pg. 
389
-
397
)
[PubMed]
7
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
, vol. 
399
 (pg. 
271
-
275
)
[PubMed]
8
Jaakkola
 
P.
Mole
 
D. R.
Tian
 
Y. M.
Wilson
 
M. I.
Gielbert
 
J.
Gaskell
 
S. J.
von Kriegsheim
 
A.
Hebestreit
 
H. F.
Mukherji
 
M.
Schofield
 
C. J.
, et al 
Targeting of HIF-α to the von Hippel–Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation
Science
2001
, vol. 
292
 (pg. 
468
-
472
)
[PubMed]
9
Ivan
 
M.
Kondo
 
K.
Yang
 
H.
Kim
 
W.
Valiando
 
J.
Ohh
 
M.
Salic
 
A.
Asara
 
J. M.
Lane
 
W. S.
Kaelin
 
W. G.
 
HIFα targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing
Science
2001
, vol. 
292
 (pg. 
464
-
468
)
[PubMed]
10
Yu
 
F.
White
 
S. B.
Zhao
 
Q.
Lee
 
F. S.
 
HIF-1α binding to VHL is regulated by stimulus-sensitive proline hydroxylation
Proc. Natl. Acad. Sci. U.S.A.
2001
, vol. 
98
 (pg. 
9630
-
9635
)
[PubMed]
11
Robinson
 
C. M.
Ohh
 
M.
 
The multifaceted von Hippel–Lindau tumour suppressor protein
FEBS Lett.
2014
 
doi:10.1016/j.febslet.2014.02.026
12
Lando
 
D.
Peet
 
D. J.
Gorman
 
J. J.
Whelan
 
D. A.
Whitelaw
 
M. L.
Bruick
 
R. K.
 
FIH-1 is an asparaginyl hydroxylase enzyme that regulates the transcriptional activity of hypoxia-inducible factor
Genes Dev.
2002
, vol. 
16
 (pg. 
1466
-
1471
)
[PubMed]
13
Schodel
 
J.
Mole
 
D. R.
Ratcliffe
 
P. J.
 
Pan-genomic binding of hypoxia-inducible transcription factors
Biol. Chem.
2013
, vol. 
394
 (pg. 
507
-
517
)
[PubMed]
14
Moniz
 
S.
Biddlestone
 
J.
Rocha
 
S.
 
Grow(2): the HIF system, energy homeostasis and the cell cycle
Histol. Histopathol.
2014
, vol. 
29
 (pg. 
589
-
600
)
[PubMed]
15
Klose
 
R. J.
Kallin
 
E. M.
Zhang
 
Y.
 
JmjC-domain-containing proteins and histone demethylation
Nat. Rev. Genet.
2006
, vol. 
7
 (pg. 
715
-
727
)
[PubMed]
16
Trewick
 
S. C.
McLaughlin
 
P. J.
Allshire
 
R. C.
 
Methylation: lost in hydroxylation?
EMBO Rep.
2005
, vol. 
6
 (pg. 
315
-
320
)
[PubMed]
17
Elkins
 
J. M.
Hewitson
 
K. S.
McNeill
 
L. A.
Seibel
 
J. F.
Schlemminger
 
I.
Pugh
 
C. W.
Ratcliffe
 
P. J.
Schofield
 
C. J.
 
Structure of factor-inhibiting hypoxia-inducible factor (HIF) reveals mechanism of oxidative modification of HIF-1α
J. Biol. Chem.
2003
, vol. 
278
 (pg. 
1802
-
1806
)
[PubMed]
18
Owen-Hughes
 
T.
Gkikopoulos
 
T.
 
Making sense of transcribing chromatin
Curr. Opin. Cell Biol.
2012
, vol. 
24
 (pg. 
296
-
304
)
[PubMed]
19
Swygert
 
S. G.
Peterson
 
C. L.
 
Chromatin dynamics: interplay between remodeling enzymes and histone modifications
Biochim. Biophys. Acta
2014
, vol. 
1839
 (pg. 
728
-
736
)
20
Lleres
 
D.
James
 
J.
Swift
 
S.
Norman
 
D. G.
Lamond
 
A. I.
 
Quantitative analysis of chromatin compaction in living cells using FLIM-FRET
J. Cell Biol.
2009
, vol. 
187
 (pg. 
481
-
496
)
[PubMed]
21
Narlikar
 
G. J.
Sundaramoorthy
 
R.
Owen-Hughes
 
T.
 
Mechanisms and functions of ATP-dependent chromatin-remodeling enzymes
Cell
2013
, vol. 
154
 (pg. 
490
-
503
)
[PubMed]
22
Zentner
 
G. E.
Henikoff
 
S.
 
Regulation of nucleosome dynamics by histone modifications
Nat. Struct. Mol. Biol.
2013
, vol. 
20
 (pg. 
259
-
266
)
[PubMed]
23
Weber
 
C. M.
Henikoff
 
S.
 
Histone variants: dynamic punctuation in transcription
Genes Dev.
2014
, vol. 
28
 (pg. 
672
-
682
)
[PubMed]
24
Thinnes
 
C. C.
England
 
K. S.
Kawamura
 
A.
Chowdhury
 
R.
Schofield
 
C. J.
Hopkinson
 
R. J.
 
Targeting histone lysine demethylases–progress, challenges, and the future
Biochim. Biophys. Acta
2014
 
doi:10.1016/j.bbagrm.2014.05.009
25
Wozniak
 
G. G.
Strahl
 
B. D.
 
Hitting the ‘mark’: interpreting lysine methylation in the context of active transcription
Biochim. Biophys. Acta
2014
 
doi:10.1016/j.bbagrm.2014.03.002
26
Rivera
 
C.
Gurard-Levin
 
Z. A.
Almouzni
 
G.
Loyola
 
A.
 
Histone lysine methylation and chromatin replication
Biochim. Biophys. Acta
2014
 
doi:10.1016/j.bbagrm.2014.03.009
27
Kooistra
 
S. M.
Helin
 
K.
 
Molecular mechanisms and potential functions of histone demethylases
Nat. Rev. Mol. Cell Biol.
2012
, vol. 
13
 (pg. 
297
-
311
)
[PubMed]
28
Clifton
 
I. J.
McDonough
 
M. A.
Ehrismann
 
D.
Kershaw
 
N. J.
Granatino
 
N.
Schofield
 
C. J.
 
Structural studies on 2-oxoglutarate oxygenases and related double-stranded β-helix fold proteins
J. Inorg. Biochem.
2006
, vol. 
100
 (pg. 
644
-
669
)
[PubMed]
29
Couture
 
J. F.
Collazo
 
E.
Ortiz-Tello
 
P. A.
Brunzelle
 
J. S.
Trievel
 
R. C.
 
Specificity and mechanism of JMJD2A, a trimethyllysine-specific histone demethylase
Nat. Struct. Mol. Biol.
2007
, vol. 
14
 (pg. 
689
-
695
)
[PubMed]
30
Aik
 
W.
McDonough
 
M. A.
Thalhammer
 
A.
Chowdhury
 
R.
Schofield
 
C. J.
 
Role of the jelly-roll fold in substrate binding by 2-oxoglutarate oxygenases
Curr. Opin. Struct. Biol.
2012
, vol. 
22
 (pg. 
691
-
700
)
[PubMed]
31
Tsukada
 
Y.-i.
Fang
 
J.
Erdjument-Bromage
 
H.
Warren
 
M. E.
Borchers
 
C. H.
Tempst
 
P.
Zhang
 
Y.
 
Histone demethylation by a family of JmjC domain-containing proteins
Nature
2006
, vol. 
439
 (pg. 
811
-
816
)
[PubMed]
32
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
, vol. 
279
 (pg. 
9899
-
9904
)
[PubMed]
33
Sanchez-Fernandez
 
E. M.
Tarhonskaya
 
H.
Al-Qahtani
 
K.
Hopkinson
 
R. J.
McCullagh
 
J. S.
Schofield
 
C. J.
Flashman
 
E.
 
Investigations on the oxygen dependence of a 2-oxoglutarate histone demethylase
Biochem. J.
2013
, vol. 
449
 (pg. 
491
-
496
)
[PubMed]
34
Melvin
 
A.
Rocha
 
S.
 
Chromatin as an oxygen sensor and active player in the hypoxia response
Cell Signal.
2012
, vol. 
24
 (pg. 
35
-
43
)
[PubMed]
35
Beyer
 
S.
Kristensen
 
M. M.
Jensen
 
K. S.
Johansen
 
J. V.
Staller
 
P.
 
The histone demethylases JMJD1A and JMJD2B are transcriptional targets of hypoxia-inducible factor HIF
J. Biol. Chem.
2008
, vol. 
283
 (pg. 
36542
-
36552
)
[PubMed]
36
Pollard
 
P.
Loenarz
 
C.
Mole
 
D.
McDonough
 
M.
Gleadle
 
J.
Schofield
 
C.
Ratcliffe
 
P.
 
Regulation of Jumonji-domain-containing histone demethylases by hypoxia-inducible factor (HIF)-1α
Biochem. J.
2008
, vol. 
416
 (pg. 
387
-
394
)
[PubMed]
37
Wellmann
 
S.
Bettkober
 
M.
Zelmer
 
A.
Seeger
 
K.
Faigle
 
M.
Eltzschig
 
H. K.
Bührer
 
C.
 
Hypoxia upregulates the histone demethylase JMJD1A via HIF-1
Biochem. Biophys. Res. Commun.
2008
, vol. 
372
 (pg. 
892
-
897
)
[PubMed]
38
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 1α enhances hypoxic gene expression and tumor growth
Mol. Cell. Biol.
2010
, vol. 
30
 (pg. 
344
-
353
)
[PubMed]
39
Niu
 
X.
Zhang
 
T.
Liao
 
L.
Zhou
 
L.
Lindner
 
D. J.
Zhou
 
M.
Rini
 
B.
Yan
 
Q.
Yang
 
H.
 
The von Hippel–Lindau tumor suppressor protein regulates gene expression and tumor growth through histone demethylase JARID1C
Oncogene
2012
, vol. 
31
 (pg. 
776
-
786
)
[PubMed]
40
Lee
 
H. Y.
Choi
 
K.
Oh
 
H.
Park
 
Y. K.
Park
 
H.
 
HIF-1-dependent induction of Jumonji domain-containing protein (JMJD) 3 under hypoxic conditions
Mol. Cells
2014
, vol. 
37
 (pg. 
43
-
50
)
[PubMed]
41
Black
 
J. C.
Van Rechem
 
C.
Whetstine
 
J. R.
 
Histone lysine methylation dynamics: establishment, regulation, and biological impact
Mol. Cell
2012
, vol. 
48
 (pg. 
491
-
507
)
[PubMed]
42
Shi
 
Y.
Whetstine
 
J. R.
 
Dynamic regulation of histone lysine methylation by demethylases
Mol. Cell
2007
, vol. 
25
 (pg. 
1
-
14
)
[PubMed]
43
Curtis
 
B. J.
Zraly
 
C. B.
Marenda
 
D. R.
Dingwall
 
A. K.
 
Histone lysine demethylases function as co-repressors of SWI/SNF remodeling activities during Drosophila wing development
Dev. Biol.
2011
, vol. 
350
 (pg. 
534
-
547
)
[PubMed]
44
Tie
 
F.
Banerjee
 
R.
Conrad
 
P. A.
Scacheri
 
P. C.
Harte
 
P. J.
 
Histone demethylase UTX and chromatin remodeler BRM bind directly to CBP and modulate acetylation of histone H3 lysine 27
Mol. Cell. Biol.
2012
, vol. 
32
 (pg. 
2323
-
2334
)
[PubMed]
45
Schnetz
 
M. P.
Bartels
 
C. F.
Shastri
 
K.
Balasubramanian
 
D.
Zentner
 
G. E.
Balaji
 
R.
Zhang
 
X.
Song
 
L.
Wang
 
Z.
Laframboise
 
T.
, et al 
Genomic distribution of CHD7 on chromatin tracks H3K4 methylation patterns
Genome Res.
2009
, vol. 
19
 (pg. 
590
-
601
)
[PubMed]
46
Menon
 
T.
Yates
 
J. A.
Bochar
 
D. A.
 
Regulation of androgen-responsive transcription by the chromatin remodeling factor CHD8
Mol. Endocrinol.
2010
, vol. 
24
 (pg. 
1165
-
1174
)
[PubMed]
47
Yates
 
J. A.
Menon
 
T.
Thompson
 
B. A.
Bochar
 
D. A.
 
Regulation of HOXA2 gene expression by the ATP-dependent chromatin remodeling enzyme CHD8
FEBS Lett.
2010
, vol. 
584
 (pg. 
689
-
693
)
[PubMed]
48
Maltby
 
V. E.
Martin
 
B. J.
Schulze
 
J. M.
Johnson
 
I.
Hentrich
 
T.
Sharma
 
A.
Kobor
 
M. S.
Howe
 
L.
 
Histone H3 lysine 36 methylation targets the Isw1b remodeling complex to chromatin
Mol. Cell. Biol.
2012
, vol. 
32
 (pg. 
3479
-
3485
)
[PubMed]
49
Musselman
 
C. A.
Mansfield
 
R. E.
Garske
 
A. L.
Davrazou
 
F.
Kwan
 
A. H.
Oliver
 
S. S.
O’Leary
 
H.
Denu
 
J. M.
Mackay
 
J. P.
Kutateladze
 
T. G.
 
Binding of the CHD4 PHD2 finger to histone H3 is modulated by covalent modifications
Biochem. J.
2009
, vol. 
423
 (pg. 
179
-
187
)
[PubMed]
50
Allen
 
H. F.
Wade
 
P. A.
Kutateladze
 
T. G.
 
The NuRD architecture
Cell. Mol. Life Sci.
2013
, vol. 
70
 (pg. 
3513
-
3524
)
[PubMed]
51
Wang
 
Y.
Zhang
 
H.
Chen
 
Y.
Sun
 
Y.
Yang
 
F.
Yu
 
W.
Liang
 
J.
Sun
 
L.
Yang
 
X.
Shi
 
L.
, et al 
LSD1 is a subunit of the NuRD complex and targets the metastasis programs in breast cancer
Cell
2009
, vol. 
138
 (pg. 
660
-
672
)
[PubMed]
52
Kenneth
 
N. S.
Mudie
 
S.
van Uden
 
P.
Rocha
 
S.
 
SWI/SNF regulates the cellular response to hypoxia
J. Biol. Chem.
2009
, vol. 
284
 (pg. 
4123
-
4131
)
[PubMed]
53
Melvin
 
A.
Mudie
 
S.
Rocha
 
S.
 
The chromatin remodeler ISWI regulates the cellular response to hypoxia: role of FIH
Mol. Biol. Cell
2011
, vol. 
22
 (pg. 
4171
-
4181
)
[PubMed]
54
Lee
 
J. S.
Kim
 
Y.
Bhin
 
J.
Shin
 
H. J.
Nam
 
H. J.
Lee
 
S. H.
Yoon
 
J. B.
Binda
 
O.
Gozani
 
O.
Hwang
 
D.
Baek
 
S. H.
 
Hypoxia-induced methylation of a pontin chromatin remodeling factor
Proc. Natl. Acad. Sci. U.S.A.
2011
, vol. 
108
 (pg. 
13510
-
13515
)
[PubMed]
55
Lee
 
J. S.
Kim
 
Y.
Kim
 
I. S.
Kim
 
B.
Choi
 
H. J.
Lee
 
J. M.
Shin
 
H. J.
Kim
 
J. H.
Kim
 
J. Y.
Seo
 
S. B.
, et al 
Negative regulation of hypoxic responses via induced Reptin methylation
Mol. Cell
2010
, vol. 
39
 (pg. 
71
-
85
)
[PubMed]
56
Johnson
 
A. B.
Denko
 
N.
Barton
 
M. C.
 
Hypoxia induces a novel signature of chromatin modifications and global repression of transcription
Mutat. Res.
2008
, vol. 
640
 (pg. 
174
-
179
)
[PubMed]
57
Zhou
 
X.
Sun
 
H.
Chen
 
H.
Zavadil
 
J.
Kluz
 
T.
Arita
 
A.
Costa
 
M.
 
Hypoxia induces trimethylated H3 lysine 4 by inhibition of JARID1A demethylase
Cancer Res.
2010
, vol. 
70
 (pg. 
4214
-
4221
)
[PubMed]
58
Tausendschön
 
M.
Dehne
 
N.
Brüne
 
B.
 
Hypoxia causes epigenetic gene regulation in macrophages by attenuating Jumonji histone demethylase activity
Cytokine
2011
, vol. 
53
 (pg. 
256
-
262
)
[PubMed]
59
Lan
 
F.
Bayliss
 
P. E.
Rinn
 
J. L.
Whetstine
 
J. R.
Wang
 
J. K.
Chen
 
S.
Iwase
 
S.
Alpatov
 
R.
Issaeva
 
I.
Canaani
 
E.
 
A histone H3 lysine 27 demethylase regulates animal posterior development
Nature
2007
, vol. 
449
 (pg. 
689
-
694
)
[PubMed]
60
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
, vol. 
121
 (pg. 
2447
-
2456
)
61
Okuno
 
Y.
Ohtake
 
F.
Igarashi
 
K.
Kanno
 
J.
Matsumoto
 
T.
Takada
 
I.
Kato
 
S.
Imai
 
Y.
 
Epigenetic regulation of adipogenesis by PHF2 histone demethylase
Diabetes
2013
, vol. 
62
 (pg. 
1426
-
1434
)
[PubMed]
62
Okuno
 
Y.
Inoue
 
K.
Imai
 
Y.
 
Novel insights into histone modifiers in adipogenesis
Adipocyte
2013
, vol. 
2
 (pg. 
285
-
288
)
[PubMed]
63
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
, vol. 
22
 (pg. 
25
-
37
)
[PubMed]
64
Wagner
 
K. W.
Alam
 
H.
Dhar
 
S. S.
Giri
 
U.
Li
 
N.
Wei
 
Y.
Giri
 
D.
Cascone
 
T.
Kim
 
J.-H.
Ye
 
Y.
 
KDM2A promotes lung tumorigenesis by epigenetically enhancing ERK1/2 signaling
J. Clin. Investig.
2013
, vol. 
123
 (pg. 
5231
-
5246
)
65
He
 
J.
Nguyen
 
A. T.
Zhang
 
Y.
 
KDM2b/JHDM1b, an H3K36me2-specific demethylase, is required for initiation and maintenance of acute myeloid leukemia
Blood
2011
, vol. 
117
 (pg. 
3869
-
3880
)
[PubMed]
66
Kottakis
 
F.
Polytarchou
 
C.
Foltopoulou
 
P.
Sanidas
 
I.
Kampranis
 
S. C.
Tsichlis
 
P. N.
 
FGF-2 regulates cell proliferation, migration, and angiogenesis through an NDY1/KDM2B-miR-101-EZH2 pathway
Mol. Cell
2011
, vol. 
43
 (pg. 
285
-
298
)
[PubMed]
67
Pryor
 
J. G.
Brown-Kipphut
 
B. A.
Iqbal
 
A.
Scott
 
G. A.
 
Microarray comparative genomic hybridization detection of copy number changes in desmoplastic melanoma and malignant peripheral nerve sheath tumor
Am. J. Dermatopathol.
2011
, vol. 
33
 (pg. 
780
-
785
)
[PubMed]
68
Yang
 
Z.-Q.
Imoto
 
I.
Fukuda
 
Y.
Pimkhaokham
 
A.
Shimada
 
Y.
Imamura
 
M.
Sugano
 
S.
Nakamura
 
Y.
Inazawa
 
J.
 
Identification of a novel gene, GASC1, within an amplicon at 9p23–24 frequently detected in esophageal cancer cell lines
Cancer Res.
2000
, vol. 
60
 (pg. 
4735
-
4739
)
[PubMed]
69
Ehrbrecht
 
A.
Müller
 
U.
Wolter
 
M.
Hoischen
 
A.
Koch
 
A.
Radlwimmer
 
B.
Actor
 
B.
Mincheva
 
A.
Pietsch
 
T.
Lichter
 
P.
 
Comprehensive genomic analysis of desmoplastic medulloblastomas: identification of novel amplified genes and separate evaluation of the different histological components
J. Pathol.
2006
, vol. 
208
 (pg. 
554
-
563
)
[PubMed]
70
Vinatzer
 
U.
Gollinger
 
M.
Müllauer
 
L.
Raderer
 
M.
Chott
 
A.
Streubel
 
B.
 
Mucosa-associated lymphoid tissue lymphoma: novel translocations including rearrangements of ODZ2, JMJD2C, and CNN3
Clin. Cancer Res.
2008
, vol. 
14
 (pg. 
6426
-
6431
)
[PubMed]
71
Hayami
 
S.
Yoshimatsu
 
M.
Veerakumarasivam
 
A.
Unoki
 
M.
Iwai
 
Y.
Tsunoda
 
T.
Field
 
H. I.
Kelly
 
J. D.
Neal
 
D. E.
Yamaue
 
H.
 
Overexpression of the JmjC histone demethylase KDM5B in human carcinogenesis: involvement in the proliferation of cancer cells through the E2F/RB pathway
Mol. Cancer
2010
, vol. 
9
 pg. 
10.1186
 
72
Dalvai
 
M.
Bystricky
 
K.
 
The role of histone modifications and variants in regulating gene expression in breast cancer
J. Mammary Gland Biol. Neoplasia
2010
, vol. 
15
 (pg. 
19
-
33
)
[PubMed]
73
Barrett
 
A.
Madsen
 
B.
Copier
 
J.
Lu
 
P. J.
Cooper
 
L.
Scibetta
 
A. G.
Burchell
 
J.
Taylor-Papadimitriou
 
J.
 
PLU-1 nuclear protein, which is upregulated in breast cancer, shows restricted expression in normal human adult tissues: a new cancer/testis antigen?
Int. J. Cancer
2002
, vol. 
101
 (pg. 
581
-
588
)
[PubMed]
74
Lu
 
T.
Jackson
 
M. W.
Wang
 
B.
Yang
 
M.
Chance
 
M. R.
Miyagi
 
M.
Gudkov
 
A. V.
Stark
 
G. R.
 
Regulation of NF-κB by NSD1/FBXL11-dependent reversible lysine methylation of p65
Proc. Natl. Acad. Sci. U.S.A.
2010
, vol. 
107
 (pg. 
46
-
51
)
[PubMed]
75
Frescas
 
D.
Guardavaccaro
 
D.
Bassermann
 
F.
Koyama-Nasu
 
R.
Pagano
 
M.
 
JHDM1B/FBXL10 is a nucleolar protein that represses transcription of ribosomal RNA genes
Nature
2007
, vol. 
450
 (pg. 
309
-
313
)
[PubMed]
76
Tzatsos
 
A.
Pfau
 
R.
Kampranis
 
S. C.
Tsichlis
 
P. N.
 
Ndy1/KDM2B immortalizes mouse embryonic fibroblasts by repressing the Ink4a/Arf locus
Proc. Natl. Acad. Sci. U.S.A.
2009
, vol. 
106
 (pg. 
2641
-
2646
)
[PubMed]
77
Pfau
 
R.
Tzatsos
 
A.
Kampranis
 
S. C.
Serebrennikova
 
O. B.
Bear
 
S. E.
Tsichlis
 
P. N.
 
Members of a family of JmjC domain-containing oncoproteins immortalize embryonic fibroblasts via a JmjC domain-dependent process
Proc. Natl. Acad. Sci. U.S.A.
2008
, vol. 
105
 (pg. 
1907
-
1912
)
[PubMed]
78
Suzuki
 
T.
Minehata
 
K. i.
Akagi
 
K.
Jenkins
 
N. A.
Copeland
 
N. G.
 
Tumor suppressor gene identification using retroviral insertional mutagenesis in Blm-deficient mice
EMBO J.
2006
, vol. 
25
 (pg. 
3422
-
3431
)
[PubMed]
79
Frescas
 
D.
Guardavaccaro
 
D.
Kuchay
 
S. M.
Kato
 
H.
Poleshko
 
A.
Basrur
 
V.
Elenitoba-Johnson
 
K. S.
Katz
 
R. A.
Pagano
 
M.
 
KDM2A represses transcription of centromeric satellite repeats and maintains the heterochromatic state
Cell Cycle
2008
, vol. 
7
 (pg. 
3539
-
3547
)
[PubMed]
80
Kauffman
 
E. C.
Robinson
 
B. D.
Downes
 
M. J.
Powell
 
L. G.
Lee
 
M. M.
Scherr
 
D. S.
Gudas
 
L. J.
Mongan
 
N. P.
 
Role of androgen receptor and associated lysine – demethylase coregulators, LSD1 and JMJD2A, in localized and advanced human bladder cancer
Mol. Carcinog.
2011
, vol. 
50
 (pg. 
931
-
944
)
[PubMed]
81
Patani
 
N.
Jiang
 
W. G.
Newbold
 
R. F.
Mokbel
 
K.
 
Histone-modifier gene expression profiles are associated with pathological and clinical outcomes in human breast cancer
Anticancer Res.
2011
, vol. 
31
 (pg. 
4115
-
4125
)
[PubMed]
82
Shi
 
L.
Sun
 
L.
Li
 
Q.
Liang
 
J.
Yu
 
W.
Yi
 
X.
Yang
 
X.
Li
 
Y.
Han
 
X.
Zhang
 
Y.
 
Histone demethylase JMJD2B coordinates H3K4/H3K9 methylation and promotes hormonally responsive breast carcinogenesis
Proc. Natl. Acad. Sci. U.S.A.
2011
, vol. 
108
 (pg. 
7541
-
7546
)
[PubMed]
83
Liu
 
G.
Bollig-Fischer
 
A.
Kreike
 
B.
van de Vijver
 
M. J.
Abrams
 
J.
Ethier
 
S. P.
Yang
 
Z.-Q.
 
Genomic amplification and oncogenic properties of the GASC1 histone demethylase gene in breast cancer
Oncogene
2009
, vol. 
28
 (pg. 
4491
-
4500
)
[PubMed]
84
Cloos
 
P. A.
Christensen
 
J.
Agger
 
K.
Maiolica
 
A.
Rappsilber
 
J.
Antal
 
T.
Hansen
 
K. H.
Helin
 
K.
 
The putative oncogene GASC1 demethylates tri-and dimethylated lysine 9 on histone H3
Nature
2006
, vol. 
442
 (pg. 
307
-
311
)
[PubMed]
85
Wissmann
 
M.
Yin
 
N.
Müller
 
J. M.
Greschik
 
H.
Fodor
 
B. D.
Jenuwein
 
T.
Vogler
 
C.
Schneider
 
R.
Günther
 
T.
Buettner
 
R.
 
Cooperative demethylation by JMJD2C and LSD1 promotes androgen receptor-dependent gene expression
Nat. Cell Biol.
2007
, vol. 
9
 (pg. 
347
-
353
)
[PubMed]
86
Teng
 
Y. C.
Lee
 
C. F.
Li
 
Y. S.
Chen
 
Y. R.
Hsiao
 
P. W.
Chan
 
M. Y.
Lin
 
F. M.
Huang
 
H. D.
Chen
 
Y. T.
Jeng
 
Y. M.
, et al 
Histone demethylase RBP2 promotes lung tumorigenesis and cancer metastasis
Cancer Res.
2013
, vol. 
73
 (pg. 
4711
-
4721
)
[PubMed]
87
Wang
 
G. G.
Song
 
J.
Wang
 
Z.
Dormann
 
H. L.
Casadio
 
F.
Li
 
H.
Luo
 
J. L.
Patel
 
D. J.
Allis
 
C. D.
 
Haematopoietic malignancies caused by dysregulation of a chromatin-binding PHD finger
Nature
2009
, vol. 
459
 (pg. 
847
-
851
)
[PubMed]
88
Scibetta
 
A. G.
Santangelo
 
S.
Coleman
 
J.
Hall
 
D.
Chaplin
 
T.
Copier
 
J.
Catchpole
 
S.
Burchell
 
J.
Taylor-Papadimitriou
 
J.
 
Functional analysis of the transcription repressor PLU-1/JARID1B
Mol. Cell. Biol.
2007
, vol. 
27
 (pg. 
7220
-
7235
)
[PubMed]
89
Yamane
 
K.
Tateishi
 
K.
Klose
 
R. J.
Fang
 
J.
Fabrizio
 
L. A.
Erdjument-Bromage
 
H.
Taylor-Papadimitriou
 
J.
Tempst
 
P.
Zhang
 
Y.
 
PLU-1 is an H3K4 demethylase involved in transcriptional repression and breast cancer cell proliferation
Mol. Cell
2007
, vol. 
25
 (pg. 
801
-
812
)
[PubMed]
90
Iwase
 
S.
Lan
 
F.
Bayliss
 
P.
de la Torre-Ubieta
 
L.
Huarte
 
M.
Qi
 
H. H.
Whetstine
 
J. R.
Bonni
 
A.
Roberts
 
T. M.
Shi
 
Y.
 
The X-linked mental retardation gene SMCX/JARID1C defines a family of histone H3 lysine 4 demethylases
Cell
2007
, vol. 
128
 (pg. 
1077
-
1088
)
[PubMed]
91
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. Genetics
2008
, vol. 
45
 (pg. 
787
-
793
)
92
Jensen
 
L. R.
Amende
 
M.
Gurok
 
U.
Moser
 
B.
Gimmel
 
V.
Tzschach
 
A.
Janecke
 
A. R.
Tariverdian
 
G.
Chelly
 
J.
Fryns
 
J.-P.
 
Mutations in the JARID1C gene, which is involved in transcriptional regulation and chromatin remodeling, cause X-linked mental retardation
Am. J. Hum. Genet.
2005
, vol. 
76
 (pg. 
227
-
236
)
[PubMed]
93
Santos
 
C.
Rodriguez-Revenga
 
L.
Madrigal
 
I.
Badenas
 
C.
Pineda
 
M.
Milà
 
M.
 
A novel mutation in JARID1C gene associated with mental retardation
Eur. J. Hum. Genet.
2006
, vol. 
14
 (pg. 
583
-
586
)
[PubMed]
94
Tzschach
 
A.
Lenzner
 
S.
Moser
 
B.
Reinhardt
 
R.
Chelly
 
J.
Fryns
 
J. P.
Kleefstra
 
T.
Raynaud
 
M.
Turner
 
G.
Ropers
 
H. H.
 
Novel JARID1C/SMCX mutations in patients with X-linked mental retardation
Hum. Mutat.
2006
, vol. 
27
 (pg. 
389
-
389
)
[PubMed]
95
Tahiliani
 
M.
Mei
 
P.
Fang
 
R.
Leonor
 
T.
Rutenberg
 
M.
Shimizu
 
F.
Li
 
J.
Rao
 
A.
Shi
 
Y.
 
The histone H3K4 demethylase SMCX links REST target genes to X-linked mental retardation
Nature
2007
, vol. 
447
 (pg. 
601
-
605
)
[PubMed]
96
Laumonnier
 
F.
Holbert
 
S.
Ronce
 
N.
Faravelli
 
F.
Lenzner
 
S.
Schwartz
 
C.
Lespinasse
 
J.
Van Esch
 
H.
Lacombe
 
D.
Goizet
 
C.
 
Mutations in PHF8 are associated with X linked mental retardation and cleft lip/cleft palate
J. Med. Genetics
2005
, vol. 
42
 (pg. 
780
-
786
)
97
Kleine-Kohlbrecher
 
D.
Christensen
 
J.
Vandamme
 
J.
Abarrategui
 
I.
Bak
 
M.
Tommerup
 
N.
Shi
 
X.
Gozani
 
O.
Rappsilber
 
J.
Salcini
 
A. E.
 
A functional link between the histone demethylase PHF8 and the transcription factor ZNF711 in X-linked mental retardation
Mol. Cell
2010
, vol. 
38
 (pg. 
165
-
178
)
[PubMed]
98
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.
2014
 
doi:10.1111/cge.12363
99
Robinson
 
G.
Parker
 
M.
Kranenburg
 
T. A.
Lu
 
C.
Chen
 
X.
Ding
 
L.
Phoenix
 
T. N.
Hedlund
 
E.
Wei
 
L.
Zhu
 
X.
, et al 
Novel mutations target distinct subgroups of medulloblastoma
Nature
2012
, vol. 
488
 (pg. 
43
-
48
)
[PubMed]
100
Grasso
 
C. S.
Wu
 
Y. M.
Robinson
 
D. R.
Cao
 
X.
Dhanasekaran
 
S. M.
Khan
 
A. P.
Quist
 
M. J.
Jing
 
X.
Lonigro
 
R. J.
Brenner
 
J. C.
, et al 
The mutational landscape of lethal castration-resistant prostate cancer
Nature
2012
, vol. 
487
 (pg. 
239
-
243
)
[PubMed]
101
Jones
 
D. T.
Jager
 
N.
Kool
 
M.
Zichner
 
T.
Hutter
 
B.
Sultan
 
M.
Cho
 
Y. J.
Pugh
 
T. J.
Hovestadt
 
V.
Stutz
 
A. M.
, et al 
Dissecting the genomic complexity underlying medulloblastoma
Nature
2012
, vol. 
488
 (pg. 
100
-
105
)
[PubMed]
102
Agger
 
K.
Cloos
 
P. A.
Rudkjaer
 
L.
Williams
 
K.
Andersen
 
G.
Christensen
 
J.
Helin
 
K.
 
The H3K27me3 demethylase JMJD3 contributes to the activation of the INK4A-ARF locus in response to oncogene- and stress-induced senescence
Genes Dev.
2009
, vol. 
23
 (pg. 
1171
-
1176
)
[PubMed]
103
Ene
 
C. I.
Edwards
 
L.
Riddick
 
G.
Baysan
 
M.
Woolard
 
K.
Kotliarova
 
S.
Lai
 
C.
Belova
 
G.
Cam
 
M.
Walling
 
J.
, et al 
Histone demethylase Jumonji D3 (JMJD3) as a tumor suppressor by regulating p53 protein nuclear stabilization
PLoS ONE
2012
, vol. 
7
 pg. 
e51407
 
[PubMed]
104
Kruidenier
 
L.
Chung
 
C. W.
Cheng
 
Z.
Liddle
 
J.
Che
 
K.
Joberty
 
G.
Bantscheff
 
M.
Bountra
 
C.
Bridges
 
A.
Diallo
 
H.
, et al 
A selective jumonji H3K27 demethylase inhibitor modulates the proinflammatory macrophage response
Nature
2012
, vol. 
488
 (pg. 
404
-
408
)
[PubMed]
105
De Santa
 
F.
Totaro
 
M. G.
Prosperini
 
E.
Notarbartolo
 
S.
Testa
 
G.
Natoli
 
G.
 
The histone H3 lysine-27 demethylase Jmjd3 links inflammation to inhibition of polycomb-mediated gene silencing
Cell
2007
, vol. 
130
 (pg. 
1083
-
1094
)
[PubMed]
106
De Santa
 
F.
Narang
 
V.
Yap
 
Z. H.
Tusi
 
B. K.
Burgold
 
T.
Austenaa
 
L.
Bucci
 
G.
Caganova
 
M.
Notarbartolo
 
S.
Casola
 
S.
, et al 
Jmjd3 contributes to the control of gene expression in LPS-activated macrophages
EMBO J.
2009
, vol. 
28
 (pg. 
3341
-
3352
)
[PubMed]
107
Schiller
 
R.
Scozzafava
 
G.
Tumber
 
A.
Wickens
 
J. R.
Bush
 
J. T.
Rai
 
G.
Lejeune
 
C.
Choi
 
H.
Yeh
 
T. L.
Chan
 
M. C.
, et al 
A cell-permeable ester derivative of the JmjC histone demethylase inhibitor IOX1
ChemMedChem
2014
, vol. 
9
 (pg. 
566
-
571
)
[PubMed]
108
Rotili
 
D.
Tomassi
 
S.
Conte
 
M.
Benedetti
 
R.
Tortorici
 
M.
Ciossani
 
G.
Valente
 
S.
Marrocco
 
B.
Labella
 
D.
Novellino
 
E.
, et al 
Pan-histone demethylase inhibitors simultaneously targeting Jumonji C and lysine-specific demethylases display high anticancer activities
J. Med. Chem.
2014
, vol. 
57
 (pg. 
42
-
55
)
[PubMed]
109
Sayegh
 
J.
Cao
 
J.
Zou
 
M. R.
Morales
 
A.
Blair
 
L. P.
Norcia
 
M.
Hoyer
 
D.
Tackett
 
A. J.
Merkel
 
J. S.
Yan
 
Q.
 
Identification of small molecule inhibitors of Jumonji AT-rich interactive domain 1B (JARID1B) histone demethylase by a sensitive high throughput screen
J. Biol. Chem.
2013
, vol. 
288
 (pg. 
9408
-
9417
)
[PubMed]
110
Rose
 
N. R.
Woon
 
E. C.
Tumber
 
A.
Walport
 
L. J.
Chowdhury
 
R.
Li
 
X. S.
King
 
O. N.
Lejeune
 
C.
Ng
 
S. S.
Krojer
 
T.
, et al 
Plant growth regulator daminozide is a selective inhibitor of human KDM2/7 histone demethylases
J. Med. Chem.
2012
, vol. 
55
 (pg. 
6639
-
6643
)
[PubMed]
111
Woon
 
E. C.
Tumber
 
A.
Kawamura
 
A.
Hillringhaus
 
L.
Ge
 
W.
Rose
 
N. R.
Ma
 
J. H.
Chan
 
M. C.
Walport
 
L. J.
Che
 
K. H.
, et al 
Linking of 2-oxoglutarate and substrate binding sites enables potent and highly selective inhibition of JmjC histone demethylases
Angew Chem. Int. Ed. Engl.
2012
, vol. 
51
 (pg. 
1631
-
1634
)
[PubMed]
112
Fukuda
 
T.
Tokunaga
 
A.
Sakamoto
 
R.
Yoshida
 
N.
 
Fbxl10/Kdm2b deficiency accelerates neural progenitor cell death and leads to exencephaly
Mol. Cell. Neurosci.
2011
, vol. 
46
 (pg. 
614
-
624
)
[PubMed]
113
Lagarou
 
A.
Mohd-Sarip
 
A.
Moshkin
 
Y. M.
Chalkley
 
G. E.
Bezstarosti
 
K.
Demmers
 
J. A.
Verrijzer
 
C. P.
 
dKDM2 couples histone H2A ubiquitylation to histone H3 demethylation during Polycomb group silencing
Genes Dev.
2008
, vol. 
22
 (pg. 
2799
-
2810
)
[PubMed]
114
Okada
 
Y.
Scott
 
G.
Ray
 
M. K.
Mishina
 
Y.
Zhang
 
Y.
 
Histone demethylase JHDM2A is critical for Tnp1 and Prm1 transcription and spermatogenesis
Nature
2007
, vol. 
450
 (pg. 
119
-
123
)
[PubMed]
115
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
, vol. 
14
 (pg. 
991
-
1001
)
[PubMed]
116
Tateishi
 
K.
Okada
 
Y.
Kallin
 
E. M.
Zhang
 
Y.
 
Role of Jhdm2a in regulating metabolic gene expression and obesity resistance
Nature
2009
, vol. 
458
 (pg. 
757
-
761
)
[PubMed]
117
Whetstine
 
J. R.
Nottke
 
A.
Lan
 
F.
Huarte
 
M.
Smolikov
 
S.
Chen
 
Z.
Spooner
 
E.
Li
 
E.
Zhang
 
G.
Colaiacovo
 
M.
 
Reversal of histone lysine trimethylation by the JMJD2 family of histone demethylases
Cell
2006
, vol. 
125
 (pg. 
467
-
481
)
[PubMed]
118
Lorbeck
 
M. T.
Singh
 
N.
Zervos
 
A.
Dhatta
 
M.
Lapchenko
 
M.
Yang
 
C.
Elefant
 
F.
 
The histone demethylase Dmel/Kdm4A controls genes required for life span and male-specific sex determination in Drosophila
Gene
2010
, vol. 
450
 (pg. 
8
-
17
)
[PubMed]
119
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
, vol. 
84
 (pg. 
1225
-
1234
)
120
Klose
 
R. J.
Yan
 
Q.
Tothova
 
Z.
Yamane
 
K.
Erdjument-Bromage
 
H.
Tempst
 
P.
Gilliland
 
D. G.
Zhang
 
Y.
Kaelin
 
W. G.
 
The retinoblastoma binding protein RBP2 is an H3K4 demethylase
Cell
2007
, vol. 
128
 (pg. 
889
-
900
)
[PubMed]
121
Gearhart
 
M. D.
Corcoran
 
C. M.
Wamstad
 
J. A.
Bardwell
 
V. J.
 
Polycomb group and SCF ubiquitin ligases are found in a novel BCOR complex that is recruited to BCL6 targets
Mol. Cell. Biol.
2006
, vol. 
26
 (pg. 
6880
-
6889
)
[PubMed]
122
Gildea
 
J. J.
Lopez
 
R.
Shearn
 
A.
 
A screen for new trithorax group genes identified little imaginal discs, the Drosophila melanogaster homologue of human retinoblastoma binding protein 2
Genetics
2000
, vol. 
156
 (pg. 
645
-
663
)
[PubMed]
123
Scibetta
 
A. G.
Burchell
 
J.
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
, vol. 
38
 (pg. 
1267
-
1277
)
[PubMed]
124
Cox
 
B. J.
Vollmer
 
M.
Tamplin
 
O.
Lu
 
M.
Biechele
 
S.
Gertsenstein
 
M.
Van Campenhout
 
C.
Floss
 
T.
Kühn
 
R.
Wurst
 
W.
 
Phenotypic annotation of the mouse X chromosome
Genome Res.
2010
, vol. 
20
 (pg. 
1154
-
1164
)
[PubMed]
125
Osada
 
T.
Ikegami
 
S.
Takiguchi-Hayashi
 
K.
Yamazaki
 
Y.
Katoh-Fukui
 
Y.
Higashinakagawa
 
T.
Sakaki
 
Y.
Takeuchi
 
T.
 
Increased anxiety and impaired pain response in puromycin-sensitive aminopeptidase gene-deficient mice obtained by a mouse gene-trap method
J. Neurosci.
1999
, vol. 
19
 (pg. 
6068
-
6078
)
[PubMed]
126
Motoyama
 
J.
Kitajima
 
K.
Kojima
 
M.
Kondo
 
S.
Takeuchi
 
T.
 
Organogenesis of the liver, thymus and spleen is affected in jumonji mutant mice
Mech. Dev.
1997
, vol. 
66
 (pg. 
27
-
37
)
[PubMed]
127
Takeuchi
 
T.
Kojima
 
M.
Nakajima
 
K.
Kondo
 
S.
 
Jumonji gene is essential for the neurulation and cardiac development of mouse embryos with a C3H/He background
Mech. Dev.
1999
, vol. 
86
 (pg. 
29
-
38
)
[PubMed]
128
Satoh
 
T.
Takeuchi
 
O.
Vandenbon
 
A.
Yasuda
 
K.
Tanaka
 
Y.
Kumagai
 
Y.
Miyake
 
T.
Matsushita
 
K.
Okazaki
 
T.
Saitoh
 
T.
 
The Jmjd3-Irf4 axis regulates M2 macrophage polarization and host responses against helminth infection
Nat. Immunol.
2010
, vol. 
11
 (pg. 
936
-
944
)
[PubMed]
129
Agger
 
K.
Cloos
 
P. A.
Christensen
 
J.
Pasini
 
D.
Rose
 
S.
Rappsilber
 
J.
Issaeva
 
I.
Canaani
 
E.
Salcini
 
A. E.
Helin
 
K.
 
UTX and JMJD3 are histone H3K27 demethylases involved in HOX gene regulation and development
Nature
2007
, vol. 
449
 (pg. 
731
-
734
)
[PubMed]
130
Copur
 
Ö.
Müller
 
J.
 
The histone H3-K27 demethylase Utx regulates HOX gene expression in Drosophila in a temporally restricted manner
Development
2013
, vol. 
140
 (pg. 
3478
-
3485
)
[PubMed]
131
Tsukada
 
Y.-i.
Ishitani
 
T.
Nakayama
 
K. I.
 
KDM7 is a dual demethylase for histone H3 Lys 9 and Lys 27 and functions in brain development
Genes Dev.
2010
, vol. 
24
 (pg. 
432
-
437
)
[PubMed]
132
Oh
 
S.
Janknecht
 
R.
 
Histone demethylase JMJD5 is essential for embryonic development
Biochem. Biophys. Res. Commun.
2012
, vol. 
420
 (pg. 
61
-
65
)
[PubMed]
133
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
, vol. 
302
 (pg. 
1560
-
1563
)
[PubMed]
134
Hong
 
J.-R.
Lin
 
G.-H.
Lin
 
C. J.-F.
Wang
 
W.-p.
Lee
 
C.-C.
Lin
 
T.-L.
Wu
 
J.-L.
 
Phosphatidylserine receptor is required for the engulfment of dead apoptotic cells and for normal embryonic development in zebrafish
Development
2004
, vol. 
131
 (pg. 
5417
-
5427
)
[PubMed]
135
Wang
 
X.
Wu
 
Y. C.
Fadok
 
V. A.
Lee
 
M. C.
Gengyo-Ando
 
K.
Cheng
 
L. C.
Ledwich
 
D.
Hsu
 
P. K.
Chen
 
J. Y.
Chou
 
B. K.
, et al 
Cell corpse engulfment mediated by C. elegans phosphatidylserine receptor through CED-5 and CED-12
Science
2003
, vol. 
302
 (pg. 
1563
-
1566
)
[PubMed]
136
Krieser
 
R. J.
Moore
 
F. E.
Dresnek
 
D.
Pellock
 
B. J.
Patel
 
R.
Huang
 
A.
Brachmann
 
C.
White
 
K.
 
The Drosophila homolog of the putative phosphatidylserine receptor functions to inhibit apoptosis
Development
2007
, vol. 
134
 (pg. 
2407
-
2414
)
[PubMed]

Author notes

1

These authors contributed equally to this work.

This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC-BY) (http://creativecommons.org/licenses/by/3.0/) which permits unrestricted use, distribution and reproduction in any medium, provided the original work is properly cited.