Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Jul 28.
Published in final edited form as: J Med Chem. 2022 Jul 15;65(14):9564–9579. doi: 10.1021/acs.jmedchem.2c00680

Recent advances with KDM4 inhibitors and potential applications

Qiong Wu 1, Brandon Young 2, Yan Wang 3, Andrew Davidoff 1, Zoran Rankovic 2, Jun Yang 1,4
PMCID: PMC9531573  NIHMSID: NIHMS1838857  PMID: 35838529

Abstract

The histone lysine demethylase 4 (KDM4) family plays an important role in regulating gene transcription, DNA repair, and metabolism. The dysregulation of KDM4 functions is associated with many human disorders, including cancer, obesity, and cardiovascular diseases. Selective and potent KDM4 inhibitors may help not only understanding the role of KDM4 in these disorders but also provide potential therapeutic opportunities. Here, we provide an overview of the field and discuss current status, challenges and opportunities laying ahead in the development of KDM4-based anticancer therapeutics.

1. Introduction

Histone methylation plays key roles in regulating chromatin structures, gene transcription and DNA repair. Histone methylation was initially thought to be irreversible until the discovery of lysine-specific demethylase 1 (LSD1).1 The dynamic methylation of histone lysine residues is regulated by histone methyltransferases (KMTs) (‘writers’) and demethylases (KDMs) (‘erasers’). KDMs are classified into two families based on the catalytic mechanisms. LSD1 (KDM1A) and LSD2 (KDM1B) are flavin adenine dinucleotide (FAD)-dependent amine oxidases,1, 2 while JmjC-domain containing demethylases (JmjC-KDMs), including families of KDM2-KDM8, are 2-oxoglutarate (2-OG) and Fe(II)-dependent hydroxylases. The mechanism of demethylation is a dioxygenase reaction.3 Briefly, an electron is transferred from the Fe(II) to the coordinated molecular oxygen, yielding Fe(III) and a superoxide radical that attacks the carbonyl group (C2) in the 2-OG. Decarboxylation of 2-OG ensues, producing succinate and CO2. During the split of molecular oxygen, a highly unstable Fe(IV)-oxo intermediate is produced, which extracts a proton from the methylated lysine. The resulting Fe(III) hydroxide subsequently hydroxylates the radical on the methyl group, forming a carbonyl group that spontaneously demethylates the lysines such as H3K9me3 (Figure 1A). KDMs demethylate a range of histone lysine residues, including H3K4, H3K9, H3K27, H3K36, H4K20 and H1.4K26. KDM4 specifically catalyzes the removal of di- and tri-methylated lysine 9, lysine 36 of histone 3 and H1.4K26.4, 5

Figure 1.

Figure 1.

(A) Demethylation mechanism of KDM4s. 2-OG shows in blue; H3K9me3 shows in red; (B) Protein architecture of KDM4s.

Overexpression of KDM4 proteins is prevalent in various cancers, including breast,6 prostate,7 colorectal,8 lung,9 blood,10 neuroblastoma,11 malignant pleural mesothelioma,12 and other cancers. Genetic depletion or pharmacologic inhibition of KDM4 has an anticancer effect,1125 making KDM4 an attractive target for the development of small molecule inhibitors. Several KDM4 inhibitors with different chemotypes have been reported; however, only one selective and potent KDM4 inhibitor, has entered in phase 1 clinical trial (TACH101), which is currently being evaluated for the treatment of gastrointestinal and high microsatellite instability (MSI-H) metastatic colorectal cancers.26

Here, we provide an overview in the development of KDM4 inhibitors. Firstly, we discuss the structural features and functions of the KDM4 protein family. Then, we review literature reported KDM4 inhibitors, which we classified on the basis of their targeted protein domains (e.g., catalytic domain and reader domain) and chemical structures. We also examine the impact of 2-oxoglutarate competition on the cellular activity of KDM4 inhibitors. We finish with the field outlook, discussing opportunities and challenges in the development of KDM4 inhibitors for clinical applications.

2. Structure of KDM4s

The KDM4 gene family consists of 6 members: KDM4A-F, of which KDM4E and KDM4F have been reported as protein noncoding pseudogenes.27 All KDM4 proteins contain JmjN and JmjC domains that are responsible for the demethylation activity. While the JmjC domain is the catalytic domain that is conserved across all the KDM4 members, the JmjN domain interacts with JmjC domain28 and stabilizes the KDM4 proteins (Figure 1B).29, 30 KDM4A-C also have non-catalytic domains including two Tudor domains and two plant homeodomains (PHD) domains, which are absent in KDM4D-F. Both the Tudor and PHD domains are reader domains that recognize specific histone lysine tails.31, 32 KDM4A-C catalyze the removal of methyl groups from both H3K9 and H3K36, however, KDM4D-E can only demethylate H3K9.33 At least some of the selectivity of the KDM4 subfamily residues arises from residues in or close to the catalytic domain.33Selectivity arising from the catalytic domain may be relatively more important for those members (e.g. KDM4D/E) that do not contain additional domains.33 Interestingly, when bound to KDM4A, H3K9me3 lysine side chain adopts a ‘W’–shaped conformation, whereas H3K36me3 forms ‘U’–shaped conformation.34 The Tudor domains of KDM4A may also bind to H3K23me3, H3K4me3 and H4K20me3.35, 36 While the KDM4B Tudor domains display exclusive binding to H3K23me3, the KDM4C Tudor domains bind specifically to H3K4me3.35 The biological relevance of the selectivity of KDM4 reader domains remains to be elucidated.

3. Function of KDM4s

KDM4A-C are broadly expressed in normal human tissues, whereas KDM4D protein is predominantly found in the human testes with much lower expression levels in other tissues.31 KDM4A-D are involved in a variety of physiological functions, including stem cell self-renewal,3741 hematopoietic stem cell maintenance,42 cell differentiation,43, 44 cell cycle,45, 46 DNA damage repair,38, 4751 metabolism,52, 53 and spermatogenesis.54 One study demonstrated that KDM4A promotes cardiac hypertrophy in response to hypertrophic stimuli in mice.55, 56 KDM4A was also found to regulate protein synthesis, an unexpected cytoplasmic and non–chromatin-related role.57

Dysregulation of KDM4s has been found in various diseases, including cancers. For example, KDM4A and KDM4B have been noted to cause copy gain at site-specific chromosomal loci such as 1q12 and 1q21,5860 which could be problematic for targeted therapies as amplifications (i.e., MCL1 in 1q21) are a mechanism of therapeutic resistance. KDM4A-C overexpression and amplification has been observed in several different subtypes of breast cancer. KDM4A and KDM4D were found overexpressed in 60% of breast cancers,6 whereas KDM4B overexpression was observed only in estrogen receptor (ER)-positive subtype breast cancer,61 as it plays an important role in estrogen-mediated signaling.6264 KDM4C is only amplified and overexpressed in basal-like subtypes.65 We and others have shown that KDM4A and KDM4B form complexes with ERα and activate ERα-mediated transcription.64, 66 Moreover, KDM4B and KDM4C are targets of HIF-1α, driving tumor growth.23, 64 Androgen receptor (AR) is a transcriptional factor that plays a crucial role for prostate cancer development. KDM4A-D proteins function as coactivators of AR.25, 67, 68 Similar to KDM4C that forms a complex with AR,25 KDM4A and KDM4D interact with AR and promote AR-dependent gene expression.67 KDM4B facilitates AR-related gene transcriptional activation by demethylating histone proteins at androgen-regulated chromatin sites and by stabilizing the AR protein.68 An additional oncogenic mechanism of KDM4B in prostate cancer is induction of alternative splicing of AR-V7,69 an AR isoform that lacks a ligand binding domain and plays an important role in progression of castration-resistant prostate cancer. KDM4A-C are overexpressed in colorectal tumors and are required for tumor cell growth.8 KDM4A forms a complex with P53 and depletion of KDM4A causes reduced proliferation in colorectal cancer (CRC) cells.70 KDM4A suppresses the p53 pathway and facilitates cellular transformation by cooperating with Ras in lung cancer cells.71 Depletion of KDM4A causes senescence, restraining the proliferation potential of cancer cells.71 KDM4B is involved in CRC tumorigenesis and progression by upregulating hypoxia-inducible genes.72 Recently, KDM4B has also been found to promote glucose metabolism, thus driving CRC progression.73 KDM4C regulates MALAT1 expression, which leads to the enhanced β-catenin signaling in CRC cell metastasis.74 KDM4C is overexpressed in lung cancer and correlates with poor prognosis in patients,9 and promotes radioresistance by activating TGF-β2/Smad signaling.9 Both KDM4A and KDM4C are required for the survival of acute myeloid leukemia (AML) cells. Knockdown or pharmacologic inhibition of KDM4A and KDM4C induced differentiation and inhibited proliferation of AML cells.22, 75 However, KDM4A has a different function from KDM4C in AML and knockdown of KDM4A and KDM4C led to different transcriptional changes.75 We have shown that KDM4B protein is highly expressed in MYCN-amplified neuroblastoma, and promotes tumor progression by interacting with N-Myc protein and regulating the Myc pathway.11 KDM4B regulates the proliferation, differentiation, and tumor growth in neuroblastoma models.

The biological consequences of KDM4 inhibition include the induction of cellular differentiation.11, 76 inhibition of cancer stem cell proliferation,16 cell cycle arrest and cell death, probably via repression of oncogenic MYC pathways and hormone-mediated signaling,11, 1719 and other cancer-related pathways. Due to the important functions of KDM4s in various cancers,13, 77, 78 developing small molecular inhibitors targeting KDM4s might be a promising strategy for cancer therapy.

4. KDM4 inhibitors

Given the importance of KDM4s in cancers, tremendous efforts have been made in the development of KDM4 inhibitors. As KDM4s are 2-oxoglutarate (2-OG) dependent dioxygenases, most of the small molecule inhibitors are 2-OG analogs and their derivatives that chelate the Fe2+ in the catalytic binding pocket, thereby suppressing KDM4 activity. However, most of these inhibitors are lacking selectivity and/or cell permeability.7987 Peptide-like inhibitors were also developed by mimicking the KDM4 substrate H3K9me3/K36me3 peptides and showed non-2-OG competitive properties. Some natural products metabolites of bacteria and plants have been identified as KDM4 inhibitors. While most of the KDM4 inhibitors target the JmjC catalytic domain, several groups reported inhibitors targeting the reader domains.

4.1. Metal-Chelating inhibitors

Hydroxamic acid inhibitors

Compounds 1 - 5 (Figure 2A) bearing hydroxamic acid group were found to mimic 2-OG and bind the Fe2+ and inhibit KDM activity. HDAC inhibitors that commonly contain the hydroxamic acid group, such as trichostatin A (TSA, 1) and suberoylanilide hydroxamic acid (SAHA, 2), were shown to inhibit KDM4E with IC50 values of 28.4 μM and 14 μM respectively.88

Figure 2.

Figure 2.

Structures of Hydroxamic acid inhibitors (A) and Hydrazide inhibitors (B)

Recently, we and others have identified ciclopirox (CPX, 3), an antifungal agent, as a pan-KDM inhibitor.76, 89 CPX inhibited KDM4B with an IC50 of 3.8 μM in a TR-FRET assay and showed potent anti-proliferative activity against neuroblastoma cells (IC50: 0.2 – 2.7 μM), but not normal human fibroblast HS68 cells, which suppressed tumor growth in mouse models.76

Compound 4a is a KDM4 inhibitor (IC50: 4.3 – 5.9 μM, ESI-TOF MS assay) developed through a bivalent strategy that combined substrate mimic (HDAC inhibitor MS-275) with a 2-OG cofactor mimic (hydroxamate scaffold).90 Its methyl ester prodrug 4b (Methylstat) blocked KYSE150 cell growth (GI50 = 5.1 μM) and induced H3K9me3 accumulation in an immunostaining-based assay (EC50 = 8.6 μM, KYSE150; 6.3 μM, MCF7). Methylstat was further modified by introducing a fluorescein isothiocyanate to develop a fluorescent probe (5), which was used in fluorescence polarization (FP) assays for the screening of KDM4A and KDM2A inhibitors.91, 92

Hydrazide inhibitors

Similar to hydroxamic acid containing compounds, hydrazides such as 6 and 7 (Figure 2B) also chelate the catalytic ferrous ion. Tetrazolylhydrazide 6 inhibits KDM4A with an IC50 of 46.64 μM in a formaldehyde dehydrogenase (FDH)-coupled assay and 2.38 μM in the antibody-based LANCE Ultra assay.93 Cocrystal structure of compound 6 in complex with KDM4D (PDB ID: 6ETS) showed that the hydrazide moiety coordinates to the active site Ni2+ (a Fe2+ replacement for crystallization purpose) with a binding mode similar to the 2-OG cofactor.94 Whereas the plant growth regulator daminozide is a selective KDM2/7 inhibitor, Compound 7, which has the two methyl groups removed from daminozide, exhibited selective KDM4E inhibition with IC50 values of 0.2 μM in an AlphaScreen assay.89

Pyridine – based inhibitors

2,4-pyridinedicarboxylic acid (8, PDCA) is a non-specific 2OG oxygenase inhibitor showing potent enzymatic inhibition across the JmjC-KDMs family,89, 95 as well as collagen and HIF prolyl hydroxylases.96 Like the binding mode of 2-OG, PDCA chelates the Fe2+ through its pyridine nitrogen and carbonyl oxygen in a bidental way (PDB ID: 2VD7). The pyridine scaffold is a common motif found in KDM4 inhibitors (Figure 3).

Figure 3.

Figure 3.

Structures of pyridine - based inhibitors

3-amino-4-pyridinecarboxylic acid 9a lacking the 2-carboxylate group has also been reported as KDM4 inhibitor selective over KDM6B and EGLN3 (pIC50: 6.2, 5.8 for KDM4C and KDM4D, RapidFire mass spectrometry (RFMS)).83 The x-ray crystal structure of 9a (PDB code: 5FP9) showed that the pyridine nitrogen formed a single coordination bond with the metal ion.

Based on this scaffold, GlaxoSmithKline (GSK) developed a cell penetrant KDM4/5 potent dual inhibitor 10 with an IC50 of 79 – 126 nM in a RFMS assay.83 In a follow-up study, the 3-aminopyridine-4-carboxylate scaffold was optimized to a pyrido[3,4-d]pyrimidin-4(3H)-one, which retained key interactions to the proteins, thus eliminating the carboxylate group leading to improvements in physicochemical properties.84 Compound 11 also showed cellular activity in a KDM4C/5C cell imaging assay (IC50: 5 μM and 4 μM).

Similar to 11, compound 12, featuring the same pyrido[3,4-d]pyrimidin-4(3H)-one scaffold, was found to be a dual KDM4/5 inhibitor (KDM4B IC50 = 31 nM, KDM5B IC50 = 23 nM, AlphaScreen assay). Compound 12 exhibited cell permeability and increased the methyl marks H3K9me3 and H3K4me3 in Hela cells overexpressing KDM4A and KDM5B demethylases.82 Improved affinity for KDM4A was achieved by engagement of residues E169 and V313 by compound 13, a spirocyclic analogue of 12 (KDM4A Ki = 4 nM, KDM5B Ki = 7 nM). The dramatic reduction from biochemical to cell-based activity is likely due to the inhibitor competition with 2-OG.86

QC6352 (compound 14) disclosed by a group at Celgene was based on a fragment lead 3-(methylamino)isonicotinic acid 9b (KDM4C IC50 = 1400 nM) developed using a structure-based drug design. QC6352 is a potent and selective KDM4 inhibitor (KDM4A-D IC50 = 35 nM −104 nM, LANCE TR-FRET assay), which displays strong antiproliferative effects (EC50 = 3.5 nM) against the KYSE-150 cell line but has no effect on normal fibroblast cell line IMR-90. Compound 14 upregulated the KDM4 substrates H3K36me3 in KYSE-150+KDM4C cells with an EC50 of 1.3 nM. In the breast cancer PDX models, QC6352 diminished tumor growth and reduced the tumor initiating cell populations.81

The imidazopyridine-7-carboxylate, compound 15, has been identified as a selective KDM4A-C inhibitor with >50 fold selectivity over any other 2-OG-dependent dioxygenases, including KDM4D, in a chemical proteomics assay.97

EPZ020809 (compound 16) was developed as a KDM4C inhibitor (Ki = 31 nM) by Epizyme, using a high-throughput mass spectrometry assay.98 A crystal structure of KDM4A complexed with compound 16 (PDB code: 4GD4) showed that the nitrogen of pyridine and pyrrole ring chelates the metal ion bidentally in a 2-OG-competitive pattern.

Molecular docking of a ZINC fragment library followed by fragment linking and hybrid optimization resulted in a pan-JmjC KDM inhibitor (compound 17) with IC50 value of 64 nM against KDM4C (2 μM 2-OG, TR-FRET assay).99 Compound 17 formed 1,5-bidentate coordination of iron through pyridine nitrogen and phenol oxygen (PDB code: 5A7W) and exhibited competition with 2-OG.

JIB-04 (compound 18) was identified as a KDM4/5 inhibitor (KDM4A-E IC50 = 290 – 1100 nM; KDM5A IC50 = 230 nM, ELISA assay) through cell-based screening of NCI’s diversity set library.80 JIB-04 blocked the proliferation of lung cancer, prostate cancer,80 rhabdomyosarcoma,100 and germ cell tumor cell lines,101 but not the normal cells,80and suppressed tumor growth in lung cancer, breast cancer,80 Ewing sarcoma,102 hepatocellular carcinoma103 xenograft mouse models. Interestingly, JIB-04 is not a competitive inhibitor of 2-OG, though it bears a metal chelating moiety.80 Mechanistic studies indicate that JIB-04 disrupts the binding of O2 and histone substrates in the KDM4A active site by interacting with Lys 241 and Tyr 177 through hydrogen bonding.85 Additionally, the inhibitor may chelate the metal center, thereby altering the binding ability of the co-substrate 2-OG. Interestingly, this study also showed that JIB-04 more effectively inhibits KDM4A in low-O2 environments, a further indication that the inhibitor is competitive with respect to O2.85

Structure optimization of two hits from virtual screening by Fang at al. led to the discovery of a KDM4D inhibitor (compound 19) with an IC50 of 0.41 μM.104 The crystal structure of compound 19 in complex with KDM4D (PDB code: 7DYQ) showed that pyridyl nitrogen, instead of the nitrogen of nitrile, coordinates to the metal ion.105

8-Hydroxyquinoline – based inhibitors

8-hydroxyquinoline (8HQ) derivatives (Figure 4A), which were first identified as 2-OG oxygenase inhibitors of EGLN,106 were also found to be KDM4 inhibitors. 5-carboxyl-8HQ (compound 20, IOX1) is a KDM4 inhibitor (KDM4A IC50: 1.7 μM, MALDI-TOF MS assay) discovered in a formaldehyde dehydrogenase (FDH)-coupled high-throughput screening assay.107 Structure-guide modification of IOX led to a cellular permeable KDM4E probe ML324 (compound 21) with IC50 value of 920 nM in an AlphaScreen assay, which displayed potent anti-viral activities.108 Another 8HQ derivative B3 (compound 22) is a KDM4 inhibitor (KDM4B IC50 = 10 nM, antibody-based fluorometric assay) with weak inhibition of KDM5A and LSD1. Compound 22 inhibited the growth of AR-positive cell lines and breast cancer cell lines with IC50 values of micromolar range in a MTT assay. In addition, compound 22 showed tumor growth inhibition in a PC3 xenograft mouse model.109 SD70 (compound 23), first discovered in a phenotypic screening of ligand and genotoxic stress-induced translocations in prostate cancer cells, was identified as a KDM4C (IC50 = 30 μM, antibody-based assay) inhibitor.110 SD70 blocks proliferation of prostate cancer cell CWR22Rv1 (concentrations > 5 μM) and inhibits the tumor growth in xenograft mouse model.

Figure 4.

Figure 4.

Structures of 8-Hydroxyquinoline - based inhibitors (A), Benzimidazole – based inhibitors (B), and clinically used iron chelating drugs (C)

Benzimidazole – based inhibitors

By means of an FDH-coupled high-throughput screening of the ChemBioNet library, Carter et al. identified the benzimidazole compound 24a (Figure 4B) as a KDM4 inhibitor (KDM4A-E IC50 = 4 – 8 μM).111 In a subsequent study, structure optimization resulted in compound 24b with 10-fold increased potency (KDM4E IC50 = 0.9 μM in FDH-based assay). 24b exhibited cytotoxicity against prostate cancer cell lines (GI50 = 8 μM for LnCap and DU145 cell lines) and a non-disease control cell line (GI50 = 26 μM, human prostate epithelial cells).87 A kinetics-based study revealed that compound 24b is not competitive with the 2-OG cosubstrate. Furthermore, crystallographic study showed that the benzimidazole pyrazole scaffold binds to a novel surface-exposed binding site formed by residues E118, S207, Y209, T261, K259, F279, and F114. In conclusion, benzimidazole pyrazole compounds inactivate the enzyme by removing metal ion in the active site.

Clinically used iron chelating drugs

Screening a set of clinically used iron chelator drugs resulted in identification of deferoxamine (25), deferiprone (26), and deferasirox (27a) (Figure 4C) as KDM4A inhibitors with IC50 values of 3.22 – 17.4 μM in FDH assay and 3.33 – 4.76 μM in LANCEUltra assay.112 Compounds 25 and 27a also show inhibition of KDM5A and KDM6B, highlighting the general selectivity challenge for iron chelator compounds. 27a increased the histone H3K9me3 level with an EC50 of 40.5 μM in KYSE-150 cells, and inhibited cell proliferation with GI50 of 3.3 μM and 5.5 μM in KYSE-150 cells and HL-60 cells, respectively. Further structure modification resulted in compounds 27b and 27c with improved cell permeability, which showed higher potency in terms of upregulation of histone trimethylation (EC50 3.3 μM for 27b, and 10 μM for 27c) and more potent inhibition of cancer cell growth (GI50 = 0.45 for KYSE-150 and 1.7 for HL-60 cell).112 Moreover, 27c did not display cytotoxicity against non-transformed lung-derived cell lines IMR-90 and BEAS-2B.

4.2. Nonmetal-chelating inhibitors

In a recent study, Fang et al. reported a KDM4D selective inhibitor 28 (Figure 5A) with a novel scaffold through AlphaLisa-based screening and structural optimization. 28 exhibited high selectivity over other KDM4 family members (KDM4D IC50 = 23 nM, KDM4A > 100 μM) and did not compete with 2-OG. Moreover, 28 displayed anti-proliferation and anti-migration in various CRC cell lines. Molecular docking results showed that 28 occupies the H3K9me3 peptide pocket, but not the 2-OG binding site.113

Figure 5.

Figure 5.

Structures of nonmetal-chelating inhibitors (A), metal cofactor disruptors (B), and peptide-based inhibitors (C)

Screening of a drug-like diversity library resulted in three hits (29 - 31) (Figure 5A) with novel scaffolds without metal-chelating moieties as KDM4A (Ki 0.42 – 3.02 μM) and KDM2A inhibitors.92 These compounds displayed cellular activity as evident by the induction of H3K9me3 and H3K36me2 in MiaPaCa2 cells with IC50 values 0.28 μM - 3.96 μM.

4.3. Metal cofactor disruptors

A Cys3-His Zn (II) binding site is present only in KDM4s, but not other 2-OG dioxygenases.34 This Zn ion site is close to the KDM4A catalytic binding pocket, and the Zn ion is important for the KDM4A stability. Zn-ejectors such as disulfiram analogs (32a and 32b) and the selenium-containing compound Ebselen (compound 33) (Figure 5B) can destabilize and inhibit KDM4A at a range of 3 μM −15 μM in a MALDI-TOF MS-based assay.114 Based on the structure of Ebselen, Kim et al. synthesized a benzo[b]tellurophene compound 34 as a KDM4 inhibitor with an IC50 value of 30 μM in a FDH assay.115 Compound 34 specifically increased the trimethylation of H3K9 but not the methylation of other lysine residues in HeLa cells and decreased the HeLa and LoVo cell viability at a concentration of 10 μM without affecting the normal human fibroblast BJ cells.

4.4. Peptide-based inhibitors

Several selective peptide inhibitors (Figure 5C) were developed through a peptidomimetic approach and showed better selectivity compared to the metal chelating inhibitors. Lohse et al. identified that H3(7–14)K9me3 and H3(7–10)K9me3 are the shortest peptides required for the catalytic activity of KDM4A and KDM4C respectively.116 The coupling of uracil, a known iron chelator,116 to the lysine 9 amino group of these peptides resulted in compound 35 with moderate KDM4A and KDM4C inhibition activities. Nielsen et al. synthesized three series of hybrid peptides where uracil-coupled lysine chain was substituted with C-substituted N-triazololysines.117 Compound 36 showed moderate KDM4C enzymatic activity in a TR-FRET assay (IC50 =30 μM).

The same substrate fragment H3K9me3 (7–14) was employed to develop selective inhibitors of the KDM4 subfamilies. Disulfide 37 was developed as a KDM4E binder by MS analysis through cross-linking of N-oxalyl-D-cysteine (NOC) to the peptides.118 The cocrystal structure for KDM4A in complex with 37 (PDB ID 3U4S) revealed 2-OG and histone binding sites were occupied by NOC and the short peptide moieties simultaneously.

Kawamura et al. developed a de novo macrocyclic peptide CP2 (Compound 38a) as a potent non-2OG competitive KDM4A-C inhibitor (33 nM - 42 nM, AlphaScreen assay) with high selectivity over other KDMs by a Peptides Integrated Discovery system.79 The crystal structure of KDM4A complexed with 38a (PDB code: 5LY1) reveals that 38a occupies the histone-binding groove of KDM4A, with Arg6 bound in the sub-pocket usually occupied by the H3K9/K36me3 residues. The positively charged residues on the 6-position are necessary for the inhibition activity. In their subsequent work, 38a was optimized to improve the cell permeability by replacing the NNK codons with NKN codons.119 The analogue 38b was more potent than 38a in KDM4A/C inhibition in an AlphaScreen assay (KDM4A IC50: 6 nM, KDM4C IC50 = 2.2 nM). Although these peptide inhibitors achieved selectivity and potency of KDM4s over other KDMs, overcoming the poor cell permeability and instability of these peptide inhibitors to achieve desired cellular activity is a big challenge.

4.5. Natural product inhibitors

A number of natural products such as (Compounds 39 - 46; Figure 6A), showed KDM4 inhibition activity.

Figure 6.

Figure 6.

Structures of natural product inhibitors (A) and Tudor domain inhibitors (B)

Tripartin (compound 39) was isolated from the culture broth of the Streptomyces sp. associated with a larva of the dung beetle Copris tripartitus Waterhouse.120 Tripartin is the first natural product which showed selective increase of the global H3K9me3 levels in Hela cells.120 It was initially defined as a selective KDM4 inhibitor. However, Guillade et al. evaluated the biological activity of synthesized tripartin and found that tripartin analogues did not inhibit isolated KDM4A–E enzymes though manifesting apparent cellular activities.121 Thus, tripartin may affect histone methylation status via a KDM4-indirect mechanism.

Geldanamycin (GA, compound 40a) is an antibiotic discovered in the culture filtrates of Streptomyces hygroscopicus var. geldanus var. nova.122 Geldanamycin derivatives have been identified as heat shock protein 90 (Hsp90) inhibitors.123 Using a TR-FRET based high-throughput screening of FDA-approved drugs and bioactive molecules,124 our team identified GA and its analogs 17-AAG (compound 40b), 17-DMAG (compound 40c) as pan-JmjC KDMs inhibitors (KDM4B/C IC50: 0.7 μM - 4.8 μM, KDM5A IC50 = 0.13 μM - 0.65 μM, KDM6B IC50 = 0.4 μM - 5.8 μM, Alpha Screen assay).125 17-DMAG upregulated H3K9me3 and H3K36me3 in the RH30 cell line, and delayed tumor growth in alveolar rhabdomyosarcoma xenograft mouse models.

Toxoflavin (compound 41) is a natural product first isolated as a toxin produced by Burkholderia gladioli from Lycoris aurea.126 Toxoflavin has shown a variety of activities, and reported as a fungicide,127 and antibiotic.128 Recently, toxoflavin has been discovered via virtual screening as a KDM4A inhibitor.129 A peptide-based histone trimethylation assay confirmed its KDM4A inhibitory activity with an IC50 value of 2.5 μM. Toxoflavin directly bound and stabilized KDM4A protein in an HCT-116 cellular thermal shift assay (CETSA) and inhibited cell proliferation and upregulated H3K9me2/3 levels. Letfus et al. investigated the binding modes using molecular docking of toxoflavin into KDM4A, and predicted the that 8-N and 7-CO oxygen of toxoflavin chelated to the metal ion.130

Based on the KDM4A inhibitor purpurogallin (compound 42a), a natural product isolated from nutgalls and oak bark, Souto et al. synthesized a series of purpurogallin analogs with improved enzymatic activity. The most potent compound 42b inhibited KDM4A-D with a modest IC50 value of 24.37 μM - 38.85 μM (AlphaLISA assay), upregulated H3K9/36me3 in MCF7 cells and induced cell death in several solid cancer cell lines.131 Interestingly, 42b did not inhibit the proliferation of normal mesenchymal cell line Mepr2B.

Flavonoids are the largest family of natural products to be identified as 2-OG dioxygenase inhibitors.132 High-throughput screening of a LOPAC compound library using a formaldehyde dehydrogenase-coupled assay resulted in KDM4E inhibitors, of which catechol 43 (epigallocatechin), flavonoids 44 (baicalein), and 45 (myricetin) were the predominant ones.133 The characterized aromatic ene–diol functionality endowed them the ability to chelate to Fe2+, which suggest that they may inhibit other 2-OG dioxygenases. Rose et al. also proved that catechols gallic acid (46a) and protocatechuic acid (46b) as pan-KDM inhibitors.89

4.6. Reader domain inhibitors

Tudor domains

Recent studies have shown that the Tudor domains of KDM4 family can direct and stimulate localized histone demethylation through binding to specific lysine methylation markers.35,134 Due to the distinct sequences and histone-binding preferences of Tudor domains between the KDM4 subfamily, targeting Tudor domains provides an alternative approach for developing selective KDM4 inhibitors.

Recently, a novel fragment 47 (Figure 6B) targeting the KDM4A-Tudor domain was identified through a 2D-NMR based fragment screening approach.135 Screening a library of rule of three compliant fragments identified fragment 47, which exhibited a modest affinity binding to nearly 80 μM. The co-crystal structure of KDM4A-Tudor domain in complex with 47 (PDB ID: 5VAR) shows that the 47 fits into the aromatic cage formed by the three conserved aromatic residues (F932, W967, Y973). The NanoBRET-based cellular proximity assay demonstrates 47 can inhibit H3K4Me3 binding to the Tudor domain in cells with an EC50 value of 105 μM.

Screening of a NIH Clinical Collection library using the HaloTag assay identified amiodarone as KDM5A-PHD3 inhibitor.136 Structure modification resulted in a series of amiodarone analogs with improved potency, of which compound 48a (WAG-003) and 48b (Figure 6B) showed inhibition against both PHD3 domain of KDM5A and Tudor domain of KDM4A with IC50 26 μM −72 μM in a fluorescence polarization assay.

PHD-finger domains

Like KDM-based Tudor domains, PHD domains are another class of reader domain for histone demethylases. While the function of many JmjC-KDM associated PHD-fingers remain unclear, some of them are partially characterized: KDM5 PHD domains bind to unmodified H3 peptide and activate catalytic domain-mediated removal of methyl marks from H3K4me3 peptide, suggesting that PHD domain inhibitor may allosterically stimulates the activity of the catalytic domain.32, 137 While a few papers reported KDM5 inhibitors targeting PHD domain,136, 138 no inhibitors targeting PHD domain of KDM4 has been identified.

5. 2-OG challenges

KDM4 is overexpressed in various cancers and plays a critical role in cancer development through transcriptional activation or regression of targeted genes. Being a promising target for cancer therapy, many efforts have been made to develop selective and potent KDM4 inhibitors. In the past decade, different types of inhibitors have been discovered to target different domains of KDM4, including catalytic domain and non-catalytic domain. However, most currently known inhibitors showed poor or no cellular activity, though they possess nanomolar-range inhibition in biochemical assays. While low cellular permeability may account for some of the observed decrease in cellular activity (Table 1), another potential factor could be competition between the cofactor 2-OG and inhibitors,86 especially those metal chelating inhibitors. Despite the different chemotypes of this class of inhibitors (hydroxamic acid, hydrazide, pyridine, 8-Hydroxyquinoline), it inevitably competes with the 2-OG when forming the coordination interaction with Fe2+ ion in the 2-OG binding pocket. In fact, 2-OG competition is common in all JmjC-KDM inhibitors. KDOAM-25, a selective KDM5A-D inhibitor, is a partial competitor of the 2-OG (IC50 = 27 nM at 4 μM 2-OG, IC50 = 558 nM at 300 μM 2-OG).139 A similar pattern was seen in CPI-455 and CPI-4203, which are KDM5 specific inhibitors.140 The KDM6 inhibitor GSK-J1 is competitive with 2-OG, but not the peptide substrate.141

Table 1.

Properties of cellular active metal-chelating inhibitors

Cmpd KDM4 IC50 Cell permeability Cellular activity IC50 Cell viability IC50 Other targets
3 KDM4B: 3.8 μM NB line: 0.2–2.7 μM KDM4C, PHF8, KDM6A
4a-b KDM4A/C/E: 3.4 – 5.9 μM H3K9me3:
MCF7: 6.3 μM
KYSE150: 8.6 μM
KYSE150: 5.1 μM KDM6B: 43 μM
PHF8: 10 μM
10 KDM4A/C/D/E: 40 – 100 nM ChromLogDpH7.4 : 0.1 Cell imaging: KDM4C: 7.9 μM KDM5C: 100 nM
KDM6B: 12 μM
11 KDM4A/C/D/E: 398 – 630 nM ChromLogDpH7.4 : 1.8 Cell imaging: KDM4C: 5 μM KDM5C: 63 nM
12 KDM4A: 102 nM
KDM4B: 31 nM
Caco-2: 11.8×10−6 cm/s Cell imaging: confirmed KDM5B: 23 nM
KDM5C: 65 nM
13 KDM4A: 100 nM
KDM4B: 43 nM
Caco-2: 11.6×10−6 cm/s IF KDM4A: 4.7 μM
InCELL Hunter:
KDM4B: 1.3 μM
KDM5B: 38 nM
KDM5C: 123 nM
14 KDM4A-D: 35 – 104 nM PAMPA: 51.2 nm/s H3K36me3: 1.3 nM KYSE150: 3.5 nM KDM5B: 750 nM
18 KDM4A-E: 290 – 1100 nM H358: 100 nM
A549: 250 nM
KDM5A: 230 nM
KDM6B: 855 nM
21 KDM4E: 920 nM Caco-2: 12.5×10−6 cm/s Viral IE infection inhibition: 10 μM
22 KDM4B: 10 nM KDM4A/C/D PC3: 40 nM
LNCaP, VcaP: submicromolar
23 KDM4C: 30 μM
24b Pan-KDM4 KDM4E: 900 nM LnCap, DU145: 8 μM
27c Pan-KDM KDM4A: 4.1 μM H3K9me3: 10 μM KYSE150: 0.46 μM
HL60: 1.7 μM

To overcome the 2-OG competition, developing covalent KDM4 inhibitors could be a potential strategy to enhance cellular activity, which has been proven to be feasible in KDM5 covalent inhibitors.142 Another way to overcome the 2-OG competition is to degrade KDM4 using Proteolysis Targeting Chimeras (PROTAC) technology. Unlike the traditional small molecule inhibitors, which require high levels of sustained occupancy, the event-driven catalytic mechanism of action makes PROTACs effective at very low, catalytic concentrations.143 Recently, a KDM5C PROTAC has been developed,144 suggesting that demethylases are tractable targets for proteasomal degradation.

However, the disconnect between potent biochemical activity and weak cellular activity for Fe chelating inhibitors may not be necessarily attributed to 2-OG competition. Compound 14 is a Fe chelator and possesses potent cellular activity in vitro and in vivo. Presumably in these settings, the concentrations of endogenous 2-OG are high. Furthermore, non-chelating inhibitors do not appear to show much cellular activity as well. These findings suggest this is a general challenge in targeting the KDMs family.

6. Perspectives of KDM4 inhibitors in clinical applications

Oncogenic transcription factors such as MYC, AR, ER, and PAX3-FOXO1 are drivers of multiple cancers including neuroblastoma, prostate cancer, breast cancer, and alveolar rhabdomyosarcoma. However, developing small molecular inhibitors targeting transcriptional factors is challenging due to the flat protein-protein interaction surface and absence of deep pockets presented in enzyme active sites.145, 146 KDM4 acts as a coactivator of those transcription factor and drives tumor progression, thus targeting KDM4 offers an alternative approach for the treatment of those cancers.

Since KDM4 is involved in DNA damage repair, pharmacologically inhibition of KDM4 by small molecules may enhance sensitivity to radiotherapy or chemotherapy that induces DNA damage. Indeed, JIB-04 was able to sensitize resistant triple-negative inflammatory breast cancer cells and their parental cell line SUM149 to the chemotherapeutic drugs doxorubicin and paclitaxel,147 restore leukemia cell sensitivity to cytarabine,148 inhibit growth of temozolomide-resistant glioblastoma cells,149 and overcome radioresistance of lung cancer cells.150 KDM4A plays an important role in innate immunity. Depletion of KDM4A has been shown to trigger tumor cell intrinsic immune responses and enhance anti-PD-1 therapy.15 JIB-04 also increases the IFN-γ+ NK cell ratio in the presence of IL-12.151 Thus, combining a KDM4 inhibitor with immunotherapy might be a promising strategy for cancer therapy.

While the functions of KDM4 in virus infection are unknown, a recent study has shown that JIB-04 has broad-spectrum antiviral activity and inhibits SARS-CoV-2 replication and coronavirus pathogenesis,152 indicating that KDM4 inhibition may has the potential to suppress infectious diseases. In a transgenic dilated cardiomyopathy mouse model, JIB-04 reduces heart size and prolongs mouse survival.153 Importantly, JIB-04 has no negative effect on healthy hearts.154 Thus, KDM4 inhibitors might be suitable to treat cardiovascular diseases.

TACH101, a selective, first-in-class, potent small-molecule inhibitor of KDM4, was shown to block the proliferation of multiple cancer cell lines and patient-derived organoid models, reduce tumor growth by up to 100% in animal xenograft models representing various cancer types (https://tachyontx.com/our-pipeline/tach101/). TACH101 has recently entered clinical trial in 2022, and its outcome is eagerly awaited by the researchers and clinicians in the field.

However, obstacles remain to translate KDM4 inhibitors to medical applications. First, the clinical benefit of KDM4 inhibition is still unclear and we have no answer yet to what patients would maximally benefit from KDM4 inhibitor treatment. While KDM4 isoforms have been demonstrated to play a role in multiple cancer types, their pathophysiology is poorly understood. Whether KDM4s are important in other cancer types needs to be studied. As KDM4s are epigenetic modulators whose activity is affected by oxygen levels, iron and 2-OG concentrations, phenotypic in vitro screening may not faithfully recapitulate the biological functions of KDM4s in vivo. Therefore, appropriate disease models are required to validate KDM4s as bona fide therapeutic targets in relevant cancers and other diseases. Patient stratification based on biomarkers are key to maximize the efficacy of targeted therapies. However, reports of suitable biomarkers for KDM4 inhibitors are scarce. Tachyon reported that colorectal cancer cells with deficiency of mismatch repair is more sensitive to TACH101. This is the first report of the clinical biomarker for KDM4 inhibition. A recent clinical investigation using PD-1 immunotherapy has achieved remarkable success in treating locally advanced rectal cancer with mismatch repair deficiency.155 Considering that KDM4 inhibition induces type I interferon response, it would be interesting to see if combination of TACH101 or other KDM4 inhibitors with immune checkpoint inhibitors could be a promising strategy to treat cancers with mismatch repair deficiency. In addition, cancers driven by dysregulated chimeric transcriptional factors or hormone receptors have shown dependency of KDM4 activity, suggesting potential therapeutic application of KDM4 inhibitors. More indications may be waiting to be discovered through further in-depth research of KDM4 physiology and pathophysiology, and the development of novel chemical probes and KDM4-based therapies is likely to continue attracting attention in both industry and academia.

ACKNOWLEDGMENTS

We are grateful for the support of the American Lebanese Syrian Associated Charities (ALSAC) and St. Jude Children’s Research Hospital, and would like to thank the patients, their families, and the staff at our institution. The work was supported by American Cancer Society-Research Scholar (130421-RSG-17-071-01-TBG, J.Y.), National Cancer Institute (1R01CA229739-01, J.Y.). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

ABBREVIATIONS USED

2-OG

2-oxoglutarate

8HQ

8-hydroxyquinoline

AML

acute myeloid leukemia

AR

androgen receptor

CETSA

cellular thermal shift assay

Chem-seq

chemical affinity capture and massively parallel DNA sequencing

CPX

ciclopirox

CRC

colorectal cancer

ER

estrogen receptor

FAD

flavin adenine dinucleotide

FDH

formaldehyde dehydrogenase

FP

fluorescence polarization

GA

geldanamycin

JmjC-KDM

JmjC-domain containing demethylase

KDM

histone lysine demethylase

KMT

histone methyltransferases

LSD1

lysine-specific demethylase 1

MSI-H

high microsatellite instability

NOC

N-oxalyl-D-cysteine

PDCA

pyridinedicarboxylic acid

PHD

plant homeodomains

PROTAC

proteolysis targeting chimeras

RFMS

RapidFire mass spectrometry

SAHA

suberoylanilide hydroxamic acid

TSA

trichostatin A

Biographies

Qiong Wu completed his Ph.D. in medicinal chemistry from Peking University (PKU) in 2018, after which he joined St Jude children’s research hospital as a postdoctoral fellow. Currently, he is working on the demethylase project under the supervision of Dr. Rankovic and Dr. Yang in the department of chemical biology & therapeutics and surgery department, respectively.

Brandon Young obtained his Ph.D. in organic synthesis from The Johns Hopkins University in 2001. After five years in the pharmaceutical industry, he joined St. Jude Children’s Research Hospital as a member of the High Throughput Chemistry Center (HTC). Currently, he is leader of the Medicinal Chemistry Center at SJCRH.

Yan Wang obtained her MD from Qingdao Medical College. She is an Associate Chief Physician in Department of Geriatrics and Occupational Disease at Qingdao Central Hospital. Her focus is clinical investigation and treatment of a variety of cardiovascular diseases and other critical illness.

Andrew M Davidoff obtained his MD from University of Pennsylvania. He is a board-certified pediatric surgeon, having completed residency training in general surgery at Duke University Medical Center and general and thoracic pediatric surgery at the Children’s Hospital of Philadelphia. He is the Chair of Department of Surgery at St Jude Children’s Research Hospital. His academic interests are focused on clinical and translational investigation and treatment of pediatric solid tumors and gene therapy.

Zoran Rankovic received his Ph.D. in organic chemistry from the University of Leeds, UK. That same year he joined Organon UK, and then continued his drug discovery career as medicinal chemistry director and research fellow at Schering-Plough, Merck, Eli Lilly, and most recently St. Jude Children’s Research Hospital, where he directs Chemistry Centers at the department of Chemical Biology and Therapeutics. His current research interests focus on developing small molecule protein degraders and epigenetic modulators for the treatment of pediatric cancers.

Jun Yang obtained his MBBS from Qingdao Medical College, MMed from National Institute of Pharmaceutical and Biological Products, China, PhD from the Institute of Cancer Research, UK. He is an Assistant Member of Department of Surgery at St Jude Children’s Research Hospital, USA. His research focuses on epigenetics, therapeutics, and drug discovery.

Footnotes

The authors declare no competing financial interest.

REFERENCES

  • 1.Shi Y; Lan F; Matson C; Mulligan P; Whetstine JR; Cole PA; Casero RA; Shi Y, Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 2004, 119 (7), 941–953. [DOI] [PubMed] [Google Scholar]
  • 2.Karytinos A; Forneris F; Profumo A; Ciossani G; Battaglioli E; Binda C; Mattevi A, A novel mammalian flavin-dependent histone demethylase. J. Biol. Chem 2009, 284 (26), 17775–17782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Shi Y; Whetstine JR, Dynamic regulation of histone lysine methylation by demethylases. Mol. Cell 2007, 25 (1), 1–14. [DOI] [PubMed] [Google Scholar]
  • 4.Trojer P; Zhang J; Yonezawa M; Schmidt A; Zheng H; Jenuwein T; Reinberg D, Dynamic histone H1 isotype 4 methylation and demethylation by histone lysine methyltransferase G9a/KMT1C and the Jumonji domain-containing JMJD2/KDM4 proteins. J. Biol. Chem 2009, 284 (13), 8395–8405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Whetstine JR; Nottke A; Lan F; Huarte M; Smolikov S; Chen Z; Spooner E; Li E; Zhang G; Colaiacovo M; Shi Y, Reversal of histone lysine trimethylation by the JMJD2 family of histone demethylases. Cell 2006, 125 (3), 467–481. [DOI] [PubMed] [Google Scholar]
  • 6.PATANI N; JIANG WG; NEWBOLD RF; MOKBEL K, Histone-modifier gene expression profiles are associated with pathological and clinical outcomes in human breast cancer. Anticancer Res. 2011, 31 (12), 4115–4125. [PubMed] [Google Scholar]
  • 7.Cloos PAC; Christensen J; Agger K; Maiolica A; Rappsilber J; Antal T; Hansen KH; Helin K, The putative oncogene GASC1 demethylates tri- and dimethylated lysine 9 on histone H3. Nature 2006, 442 (7100), 307–311. [DOI] [PubMed] [Google Scholar]
  • 8.Berry WL; Janknecht R, KDM4/JMJD2 histone demethylases: epigenetic regulators in cancer cells. Cancer Res. 2013, 73 (10), 2936–2942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Jie X; Fong WP; Zhou R; Zhao Y; Zhao Y; Meng R; Zhang S; Dong X; Zhang T; Yang K; Wu G; Xu S, USP9X-mediated KDM4C deubiquitination promotes lung cancer radioresistance by epigenetically inducing TGF-β2 transcription. Cell Death Differ. 2021, 28 (7), 2095–2111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Boila LD; Chatterjee SS; Banerjee D; Sengupta A, KDM6 and KDM4 histone lysine demethylases emerge as molecular therapeutic targets in human acute myeloid leukemia. Exp. Hematol 2018, 58, 44–51.e47. [DOI] [PubMed] [Google Scholar]
  • 11.Yang J; AlTahan AM; Hu D; Wang Y; Cheng P-H; Morton CL; Qu C; Nathwani AC; Shohet JM; Fotsis T; Koster J; Versteeg R; Okada H; Harris AL; Davidoff AM, The role of histone demethylase KDM4B in Myc signaling in neuroblastoma. J. Natl. Cancer Inst 2015, 107 (6), djv080. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Lapidot M; Case AE; Weisberg EL; Meng C; Walker SR; Garg S; Ni W; Podar K; Hung YP; Carrasco RD; Knott A; Gokhale PC; Sharma S; Pozhitkov A; Kulkarni P; Frank DA; Salgia R; Griffin JD; Saladi SV; Bueno R; Sattler M, Essential role of the histone lysine demethylase KDM4A in the biology of malignant pleural mesothelioma (MPM). Br. J. Cancer 2021, 125 (4), 582–592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Lee DH; Kim GW; Jeon YH; Yoo J; Lee SW; Kwon SH, Advances in histone demethylase KDM4 as cancer therapeutic targets. FASEB J. 2020, 34 (3), 3461–3484. [DOI] [PubMed] [Google Scholar]
  • 14.Bhan A; Ansari KI; Chen MY; Jandial R, Inhibition of Jumonji histone demethylases selectively suppresses HER2+ breast leptomeningeal carcinomatosis growth via inhibition of GM-CSF expression. Cancer Res. 2021, 81 (12), 3200–3214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Zhang W; Liu W; Jia L; Chen D; Chang I; Lake M; Bentolila LA; Wang CY, Targeting KDM4A epigenetically activates tumor-cell-intrinsic immunity by inducing DNA replication stress. Mol. Cell 2021, 81 (10), 2148–2165 e2149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Metzger E; Stepputtis SS; Strietz J; Preca BT; Urban S; Willmann D; Allen A; Zenk F; Iovino N; Bronsert P; Proske A; Follo M; Boerries M; Stickeler E; Xu J; Wallace MB; Stafford JA; Kanouni T; Maurer J; Schule R, KDM4 inhibition targets breast cancer stem-like cells. Cancer Res. 2017, 77 (21), 5900–5912. [DOI] [PubMed] [Google Scholar]
  • 17.Wu MJ; Chen CJ; Lin TY; Liu YY; Tseng LL; Cheng ML; Chuu CP; Tsai HK; Kuo WL; Kung HJ; Wang WC, Targeting KDM4B that coactivates c-Myc-regulated metabolism to suppress tumor growth in castration-resistant prostate cancer. Theranostics 2021, 11 (16), 7779–7796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Tang DE; Dai Y; He JX; Lin LW; Leng QX; Geng XY; Fu DX; Jiang HW; Xu SH, Targeting the KDM4B-AR-c-Myc axis promotes sensitivity to androgen receptor-targeted therapy in advanced prostate cancer. J. Pathol 2020, 252 (2), 101–113. [DOI] [PubMed] [Google Scholar]
  • 19.Qiu MT; Fan Q; Zhu Z; Kwan SY; Chen L; Chen JH; Ying ZL; Zhou Y; Gu W; Wang LH; Cheng WW; Zeng J; Wan XP; Mok SC; Wong KK; Bao W, KDM4B and KDM4A promote endometrial cancer progression by regulating androgen receptor, c-myc, and p27kip1. Oncotarget 2015, 6 (31), 31702–31720. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Wang W; Oguz G; Lee PL; Bao Y; Wang P; Terp MG; Ditzel HJ; Yu Q, KDM4B-regulated unfolded protein response as a therapeutic vulnerability in PTEN-deficient breast cancer. J. Exp. Med 2018, 215 (11), 2833–2849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Chen Y; Fang R; Yue C; Chang G; Li P; Guo Q; Wang J; Zhou A; Zhang S; Fuller GN; Shi X; Huang S, Wnt-induced stabilization of KDM4C is required for Wnt/β-catenin target gene expression and glioblastoma tumorigenesis. Cancer Res. 2020, 80 (5), 1049–1063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Cheung N; Fung TK; Zeisig BB; Holmes K; Rane JK; Mowen KA; Finn MG; Lenhard B; Chan LC; So CW, Targeting aberrant epigenetic networks mediated by PRMT1 and KDM4C in acute myeloid leukemia. Cancer Cell 2016, 29 (1), 32–48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Luo W; Chang R; Zhong J; Pandey A; Semenza GL, Histone demethylase JMJD2C is a coactivator for hypoxia-inducible factor 1 that is required for breast cancer progression. Proc. Natl. Acad. Sci. U. S. A 2012, 109 (49), E3367–E3376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Rui L; Emre NC; Kruhlak MJ; Chung HJ; Steidl C; Slack G; Wright GW; Lenz G; Ngo VN; Shaffer AL; Xu W; Zhao H; Yang Y; Lamy L; Davis RE; Xiao W; Powell J; Maloney D; Thomas CJ; Moller P; Rosenwald A; Ott G; Muller-Hermelink HK; Savage K; Connors JM; Rimsza LM; Campo E; Jaffe ES; Delabie J; Smeland EB; Weisenburger DD; Chan WC; Gascoyne RD; Levens D; Staudt LM, Cooperative epigenetic modulation by cancer amplicon genes. Cancer Cell 2010, 18 (6), 590–605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Wissmann M; Yin N; Muller JM; Greschik H; Fodor BD; Jenuwein T; Vogler C; Schneider R; Gunther T; Buettner R; Metzger E; Schule R, Cooperative demethylation by JMJD2C and LSD1 promotes androgen receptor-dependent gene expression. Nat. Cell Biol 2007, 9 (3), 347–353. [DOI] [PubMed] [Google Scholar]
  • 26.A study to assess the safety, pharmacokinetics, and antitumor activity of oral TACH101 in participants with advanced or metastatic cancer. https://ClinicalTrials.gov/show/NCT05076552. (accessed 2022-03-29).
  • 27.Katoh M; Katoh M, Identification and characterization of JMJD2 family genes in silico. Int. J. Oncol 2004, 24 (6), 1623–1628. [PubMed] [Google Scholar]
  • 28.Chen Z; Zang J; Whetstine J; Hong X; Davrazou F; Kutateladze TG; Simpson M; Mao Q; Pan C-H; Dai S; Hagman J; Hansen K; Shi Y; Zhang G, Structural insights into histone demethylation by JMJD2 family members. Cell 2006, 125 (4), 691–702. [DOI] [PubMed] [Google Scholar]
  • 29.Levin M; Stark M; Assaraf YG, The JmjN domain as a dimerization interface and a targeted inhibitor of KDM4 demethylase activity. Oncotarget 2018, 9 (24). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Huang F; Chandrasekharan MB; Chen Y-C; Bhaskara S; Hiebert SW; Sun Z-W, The JmjN domain of Jhd2 is important for its protein stability, and the plant homeodomain (PHD) finger mediates its chromatin association independent of H3K4 methylation. J. Biol. Chem 2010, 285 (32), 24548–24561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Labbé RM; Holowatyj A; Yang Z-Q, Histone lysine demethylase (KDM) subfamily 4: structures, functions and therapeutic potential. Am. J. Transl. Res 2013, 6 (1), 1–15. [PMC free article] [PubMed] [Google Scholar]
  • 32.Torres IO; Kuchenbecker KM; Nnadi CI; Fletterick RJ; Kelly MJS; Fujimori DG, Histone demethylase KDM5A is regulated by its reader domain through a positive-feedback mechanism. Nat. Commun 2015, 6 (1), 6204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Hillringhaus L; Yue WW; Rose NR; Ng SS; Gileadi C; Loenarz C; Bello SH; Bray JE; Schofield CJ; Oppermann U, Structural and evolutionary basis for the dual substrate selectivity of human KDM4 histone demethylase family. J. Biol. Chem 2011, 286 (48), 41616–41625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Ng SS; Kavanagh KL; McDonough MA; Butler D; Pilka ES; Lienard BMR; Bray JE; Savitsky P; Gileadi O; von Delft F; Rose NR; Offer J; Scheinost JC; Borowski T; Sundstrom M; Schofield CJ; Oppermann U, Crystal structures of histone demethylase JMJD2A reveal basis for substrate specificity. Nature 2007, 448 (7149), 87–91. [DOI] [PubMed] [Google Scholar]
  • 35.Su Z; Wang F; Lee J-H; Stephens KE; Papazyan R; Voronina E; Krautkramer KA; Raman A; Thorpe JJ; Boersma MD; Kuznetsov VI; Miller MD; Taverna SD; Phillips GN; Denu JM, Reader domain specificity and lysine demethylase-4 family function. Nat. Commun 2016, 7 (1), 13387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Huang Y; Fang J; Bedford MT; Zhang Y; Xu R-M, Recognition of histone H3 lysine-4 methylation by the double tudor domain of JMJD2A. Science 2006, 312 (5774), 748–751. [DOI] [PubMed] [Google Scholar]
  • 37.Pedersen MT; Kooistra SM; Radzisheuskaya A; Laugesen A; Johansen JV; Hayward DG; Nilsson J; Agger K; Helin K, Continual removal of H3K9 promoter methylation by Jmjd2 demethylases is vital for ESC self-renewal and early development. EMBO J. 2016, 35 (14), 1550–1564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Pfister SX; Ahrabi S; Zalmas LP; Sarkar S; Aymard F; Bachrati CZ; Helleday T; Legube G; La Thangue NB; Porter AC; Humphrey TC, SETD2-dependent histone H3K36 trimethylation is required for homologous recombination repair and genome stability. Cell Rep. 2014, 7 (6), 2006–2018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Das PP; Shao Z; Beyaz S; Apostolou E; Pinello L; De Los Angeles A; O’Brien K; Atsma JM; Fujiwara Y; Nguyen M; Ljuboja D; Guo G; Woo A; Yuan GC; Onder T; Daley G; Hochedlinger K; Kim J; Orkin SH, Distinct and combinatorial functions of Jmjd2b/Kdm4b and Jmjd2c/Kdm4c in mouse embryonic stem cell identity. Mol. Cell 2014, 53 (1), 32–48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Deng P; Yuan Q; Cheng Y; Li J; Liu Z; Liu Y; Li Y; Su T; Wang J; Salvo ME; Wang W; Fan G; Lyons K; Yu B; Wang CY, Loss of KDM4B exacerbates bone-fat imbalance and mesenchymal stromal cell exhaustion in skeletal aging. Cell Stem Cell 2021, 28 (6), 1057–1073 e1057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Wang LY; Hung CL; Chen YR; Yang JC; Wang J; Campbell M; Izumiya Y; Chen HW; Wang WC; Ann DK; Kung HJ, KDM4A coactivates E2F1 to regulate PDK-dependent metabolic switch between mitochondrial oxidation and glycolysis. Cell Rep. 2016, 16 (11), 3016–3027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Agger K; Nishimura K; Miyagi S; Messling JE; Rasmussen KD; Helin K, The KDM4/JMJD2 histone demethylases are required for hematopoietic stem cell maintenance. Blood 2019, 134 (14), 1154–1158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Verrier L; Escaffit F; Chailleux C; Trouche D; Vandromme M, A new isoform of the histone demethylase JMJD2A/KDM4A is required for skeletal muscle differentiation. PLoS Genet. 2011, 7 (6), e1001390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Strobl-Mazzulla PH; Sauka-Spengler T; Bronner-Fraser M, Histone demethylase JmjD2A regulates neural crest specification. Dev. Cell 2010, 19 (3), 460–468. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Black JC; Allen A; Van Rechem C; Forbes E; Longworth M; Tschop K; Rinehart C; Quiton J; Walsh R; Smallwood A; Dyson NJ; Whetstine JR, Conserved antagonism between JMJD2A/KDM4A and HP1gamma during cell cycle progression. Mol. Cell 2010, 40 (5), 736–748. [DOI] [PubMed] [Google Scholar]
  • 46.Kupershmit I; Khoury-Haddad H; Awwad SW; Guttmann-Raviv N; Ayoub N, KDM4C (GASC1) lysine demethylase is associated with mitotic chromatin and regulates chromosome segregation during mitosis. Nucleic Acids Res. 2014, 42 (10), 6168–6182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Janssen A; Colmenares SU; Lee T; Karpen GH, Timely double-strand break repair and pathway choice in pericentromeric heterochromatin depend on the histone demethylase dKDM4A. Genes Dev. 2019, 33 (1–2), 103–115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Mallette FA; Mattiroli F; Cui G; Young LC; Hendzel MJ; Mer G; Sixma TK; Richard S, RNF8- and RNF168-dependent degradation of KDM4A/JMJD2A triggers 53BP1 recruitment to DNA damage sites. EMBO J. 2012, 31 (8), 1865–1878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Sulkowski PL; Oeck S; Dow J; Economos NG; Mirfakhraie L; Liu Y; Noronha K; Bao X; Li J; Shuch BM; King MC; Bindra RS; Glazer PM, Oncometabolites suppress DNA repair by disrupting local chromatin signalling. Nature 2020, 582 (7813), 586–591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Hsieh HJ; Zhang W; Lin SH; Yang WH; Wang JZ; Shen J; Zhang Y; Lu Y; Wang H; Yu J; Mills GB; Peng G, Systems biology approach reveals a link between mTORC1 and G2/M DNA damage checkpoint recovery. Nat Commun 2018, 9 (1), 3982. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Young LC; McDonald DW; Hendzel MJ, Kdm4b histone demethylase is a DNA damage response protein and confers a survival advantage following gamma-irradiation. J. Biol. Chem 2013, 288 (29), 21376–21388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Zhao E; Ding J; Xia Y; Liu M; Ye B; Choi JH; Yan C; Dong Z; Huang S; Zha Y; Yang L; Cui H; Ding HF, KDM4C and ATF4 cooperate in transcriptional control of amino acid metabolism. Cell Rep. 2016, 14 (3), 506–519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Cheng Y; Yuan Q; Vergnes L; Rong X; Youn JY; Li J; Yu Y; Liu W; Cai H; Lin JD; Tontonoz P; Hong C; Reue K; Wang CY, KDM4B protects against obesity and metabolic dysfunction. Proc. Natl. Acad. Sci. U. S. A 2018, 115 (24), E5566–E5575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Iwamori N; Zhao M; Meistrich ML; Matzuk MM, 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 (6), 1225–1234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Zhang QJ; Chen HZ; Wang L; Liu DP; Hill JA; Liu ZP, The histone trimethyllysine demethylase JMJD2A promotes cardiac hypertrophy in response to hypertrophic stimuli in mice. J. Clin. Invest 2011, 121 (6), 2447–2456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Rosales W; Lizcano F, The histone demethylase JMJD2A modulates the induction of hypertrophy markers in iPSC-derived cardiomyocytes. Front. Genet 2018, 9, 14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Van Rechem C; Black JC; Boukhali M; Aryee MJ; Gräslund S; Haas W; Benes CH; Whetstine JR, Lysine demethylase KDM4A associates with translation machinery and regulates protein synthesis. Cancer Discov. 2015, 5 (3), 255–263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Mishra S; Van Rechem C; Pal S; Clarke TL; Chakraborty D; Mahan SD; Black JC; Murphy SE; Lawrence MS; Daniels DL; Whetstine JR, Cross-talk between lysine-modifying enzymes controls site-specific DNA amplifications. Cell 2018, 174 (4), 803–817.e816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Black JC; Atabakhsh E; Kim J; Biette KM; Van Rechem C; Ladd B; Burrowes PD; Donado C; Mattoo H; Kleinstiver BP; Song B; Andriani G; Joung JK; Iliopoulos O; Montagna C; Pillai S; Getz G; Whetstine JR, Hypoxia drives transient site-specific copy gain and drug-resistant gene expression. Genes Dev. 2015, 29 (10), 1018–1031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Black JC; Manning AL; Van Rechem C; Kim J; Ladd B; Cho J; Pineda CM; Murphy N; Daniels DL; Montagna C; Lewis PW; Glass K; Allis CD; Dyson NJ; Getz G; Whetstine JR, KDM4A lysine demethylase induces site-specific copy gain and rereplication of regions amplified in tumors. Cell 2013, 154 (3), 541–555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Ye Q; Holowatyj A; Wu J; Liu H; Zhang L; Suzuki T; Yang Z-Q, Genetic alterations of KDM4 subfamily and therapeutic effect of novel demethylase inhibitor in breast cancer. Am. J. Cancer Res 2015, 5 (4), 1519–1530. [PMC free article] [PubMed] [Google Scholar]
  • 62.Gaughan L; Stockley J; Coffey K; O’Neill D; Jones DL; Wade M; Wright J; Moore M; Tse S; Rogerson L; Robson CN, KDM4B is a master regulator of the estrogen receptor signalling cascade. Nucleic Acids Res. 2013, 41 (14), 6892–6904. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Kawazu M; Saso K; Tong KI; McQuire T; Goto K; Son DO; Wakeham A; Miyagishi M; Mak TW; 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 (3), e17830. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Yang J; Jubb AM; Pike L; Buffa FM; Turley H; Baban D; Leek R; Gatter KC; Ragoussis J; Harris AL, The histone demethylase JMJD2B is regulated by estrogen receptor alpha and hypoxia, and is a key mediator of estrogen induced growth. Cancer Res. 2010, 70 (16), 6456–6466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Liu G; Bollig-Fischer A; Kreike B; van de Vijver MJ; Abrams J; Ethier SP; Yang ZQ, Genomic amplification and oncogenic properties of the GASC1 histone demethylase gene in breast cancer. Oncogene 2009, 28 (50), 4491–4500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Berry WL; Shin S; Lightfoot SA; Janknecht R, Oncogenic features of the JMJD2A histone demethylase in breast cancer. Int. J. Oncol 2012, 41 (5), 1701–1706. [DOI] [PubMed] [Google Scholar]
  • 67.Shin S; Janknecht R, Activation of androgen receptor by histone demethylases JMJD2A and JMJD2D. Biochem. Biophys. Res. Commun 2007, 359 (3), 742–746. [DOI] [PubMed] [Google Scholar]
  • 68.Coffey K; Rogerson L; Ryan-Munden C; Alkharaif D; Stockley J; Heer R; Sahadevan K; O’Neill D; Jones D; Darby S; Staller P; Mantilla A; Gaughan L; Robson CN, The lysine demethylase, KDM4B, is a key molecule in androgen receptor signalling and turnover. Nucleic Acids Res. 2013, 41 (8), 4433–4446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Duan L; Chen Z; Lu J; Liang Y; Wang M; Roggero CM; Zhang QJ; Gao J; Fang Y; Cao J; Lu J; Zhao H; Dang A; Pong RC; Hernandez E; Chang CM; Hoang DT; Ahn JM; Xiao G; Wang RT; Yu KJ; Kapur P; Rizo J; Hsieh JT; Luo J; Liu ZP, Histone lysine demethylase KDM4B regulates the alternative splicing of the androgen receptor in response to androgen deprivation. Nucleic Acids Res. 2019, 47 (22), 11623–11636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Kim T-D; Shin S; Berry WL; Oh S; Janknecht R, The JMJD2A demethylase regulates apoptosis and proliferation in colon cancer cells. J. Cell. Biochem 2012, 113 (4), 1368–1376. [DOI] [PubMed] [Google Scholar]
  • 71.Mallette, Frédérick A; Richard S, JMJD2A promotes cellular transformation by blocking cellular senescence through transcriptional repression of the tumor suppressor CHD5. Cell Rep. 2012, 2 (5), 1233–1243. [DOI] [PubMed] [Google Scholar]
  • 72.Fu L; Chen L; Yang J; Ye T; Chen Y; Fang J, HIF-1α-induced histone demethylase JMJD2B contributes to the malignant phenotype of colorectal cancer cells via an epigenetic mechanism. Carcinogenesis 2012, 33 (9), 1664–1673. [DOI] [PubMed] [Google Scholar]
  • 73.Li H; Lan J; Wang G; Guo K; Han C; Li X; Hu J; Cao Z; Luo X, KDM4B facilitates colorectal cancer growth and glucose metabolism by stimulating TRAF6-mediated AKT activation. J. Exp. Clin. Cancer Res 2020, 39 (1), 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Wu X; Li R; Song Q; Zhang C; Jia R; Han Z; Zhou L; Sui H; Liu X; Zhu H; Yang L; Wang Y; Ji Q; Li Q, JMJD2C promotes colorectal cancer metastasis via regulating histone methylation of MALAT1 promoter and enhancing β-catenin signaling pathway. J. Exp. Clin. Cancer Res 2019, 38 (1), 435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Massett ME; Monaghan L; Patterson S; Mannion N; Bunschoten RP; Hoose A; Marmiroli S; Liskamp RMJ; Jørgensen HG; Vetrie D; Michie AM; Huang X, A KDM4A-PAF1-mediated epigenomic network is essential for acute myeloid leukemia cell self-renewal and survival. Cell Death Dis. 2021, 12 (6), 573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Yang J; Milasta S; Hu D; AlTahan AM; Interiano RB; Zhou J; Davidson J; Low J; Lin W; Bao J; Goh P; Nathwani AC; Wang R; Wang Y; Ong SS; Boyd VA; Young B; Das S; Shelat A; Wu Y; Li Z; Zheng JJ; Mishra A; Cheng Y; Qu C; Peng J; Green DR; White S; Guy RK; Chen T; Davidoff AM, Targeting histone demethylases in MYC-driven neuroblastomas with ciclopirox. Cancer Res. 2017, 77 (17), 4626–4638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Wilson C; Krieg AJ, KDM4B: a nail for every hammer? Genes 2019, 10 (2). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Baby S; Gurukkala Valapil D; Shankaraiah N, Unravelling KDM4 histone demethylase inhibitors for cancer therapy. Drug Discov. Today 2021, 26 (8), 1841–1856. [DOI] [PubMed] [Google Scholar]
  • 79.Kawamura A; Münzel M; Kojima T; Yapp C; Bhushan B; Goto Y; Tumber A; Katoh T; King ONF; Passioura T; Walport LJ; Hatch SB; Madden S; Müller S; Brennan PE; Chowdhury R; Hopkinson RJ; Suga H; Schofield CJ, Highly selective inhibition of histone demethylases by de novo macrocyclic peptides. Nat. Commun 2017, 8 (1), 14773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Wang L; Chang J; Varghese D; Dellinger M; Kumar S; Best AM; Ruiz J; Bruick R; Peña-Llopis S; Xu J; Babinski DJ; Frantz DE; Brekken RA; Quinn AM; Simeonov A; Easmon J; Martinez ED, A small molecule modulates Jumonji histone demethylase activity and selectively inhibits cancer growth. Nat. Commun 2013, 4 (1), 2035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Chen YK; Bonaldi T; Cuomo A; Del Rosario JR; Hosfield DJ; Kanouni T; Kao S; Lai C; Lobo NA; Matuszkiewicz J; McGeehan A; O’Connell SM; Shi L; Stafford JA; Stansfield RK; Veal JM; Weiss MS; Yuen NY; Wallace MB, Design of KDM4 inhibitors with antiproliferative effects in cancer models. ACS Med. Chem. Lett 2017, 8 (8), 869–874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Bavetsias V; Lanigan RM; Ruda GF; Atrash B; McLaughlin MG; Tumber A; Mok NY; Le Bihan Y-V; Dempster S; Boxall KJ; Jeganathan F; Hatch SB; Savitsky P; Velupillai S; Krojer T; England KS; Sejberg J; Thai C; Donovan A; Pal A; Scozzafava G; Bennett JM; Kawamura A; Johansson C; Szykowska A; Gileadi C; Burgess-Brown NA; von Delft F; Oppermann U; Walters Z; Shipley J; Raynaud FI; Westaway SM; Prinjha RK; Fedorov O; Burke R; Schofield CJ; Westwood IM; Bountra C; Müller S; van Montfort RLM; Brennan PE; Blagg J, 8-Substituted pyrido[3,4-d]pyrimidin-4(3H)-one derivatives as potent, cell permeable, KDM4 (JMJD2) and KDM5 (JARID1) histone lysine demethylase inhibitors. J. Med. Chem 2016, 59 (4), 1388–1409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Westaway SM; Preston AGS; Barker MD; Brown F; Brown JA; Campbell M; Chung C.-w.; Diallo H; Douault C; Drewes G; Eagle R; Gordon L; Haslam C; Hayhow TG; Humphreys PG; Joberty G; Katso R; Kruidenier L; Leveridge M; Liddle J; Mosley J; Muelbaier M; Randle R; Rioja I; Rueger A; Seal GA; Sheppard RJ; Singh O; Taylor J; Thomas P; Thomson D; Wilson DM; Lee K; Prinjha RK, Cell penetrant inhibitors of the KDM4 and KDM5 families of histone lysine demethylases. 1. 3-Amino-4-pyridine carboxylate derivatives. J. Med. Chem 2016, 59 (4), 1357–1369. [DOI] [PubMed] [Google Scholar]
  • 84.Westaway SM; Preston AGS; Barker MD; Brown F; Brown JA; Campbell M; Chung C.-w.; Drewes G; Eagle R; Garton N; Gordon L; Haslam C; Hayhow TG; Humphreys PG; Joberty G; Katso R; Kruidenier L; Leveridge M; Pemberton M; Rioja I; Seal GA; Shipley T; Singh O; Suckling CJ; Taylor J; Thomas P; Wilson DM; Lee K; Prinjha RK, Cell penetrant inhibitors of the KDM4 and KDM5 families of histone lysine demethylases. 2. Pyrido[3,4-d]pyrimidin-4(3H)-one derivatives. J. Med. Chem 2016, 59 (4), 1370–1387. [DOI] [PubMed] [Google Scholar]
  • 85.Cascella B; Lee SG; Singh S; Jez JM; Mirica LM, The small molecule JIB-04 disrupts O2 binding in the Fe-dependent histone demethylase KDM4A/JMJD2A. Chem. Commun 2017, 53 (13), 2174–2177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Le Bihan Y-V; Lanigan RM; Atrash B; McLaughlin MG; Velupillai S; Malcolm AG; England KS; Ruda GF; Mok NY; Tumber A; Tomlin K; Saville H; Shehu E; McAndrew C; Carmichael L; Bennett JM; Jeganathan F; Eve P; Donovan A; Hayes A; Wood F; Raynaud FI; Fedorov O; Brennan PE; Burke R; van Montfort RLM; Rossanese OW; Blagg J; Bavetsias V, C8-substituted pyrido[3,4-d]pyrimidin-4(3H)-ones: Studies towards the identification of potent, cell penetrant Jumonji C domain containing histone lysine demethylase 4 subfamily (KDM4) inhibitors, compound profiling in cell-based target engagement assays. Eur. J. Med. Chem 2019, 177, 316–337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Carter DM; Specker E; Małecki PH; Przygodda J; Dudaniec K; Weiss MS; Heinemann U; Nazaré M; Gohlke U, Enhanced properties of a benzimidazole benzylpyrazole lysine demethylase inhibitor: mechanism-of-action, binding site analysis, and activity in cellular models of prostate cancer. J. Med. Chem 2021, 64 (19), 14266–14282. [DOI] [PubMed] [Google Scholar]
  • 88.Rose NR; Ng SS; Mecinović J; Liénard BMR; Bello SH; Sun Z; McDonough MA; Oppermann U; Schofield CJ, Inhibitor scaffolds for 2-oxoglutarate-dependent histone lysine demethylases. J. Med. Chem 2008, 51 (22), 7053–7056. [DOI] [PubMed] [Google Scholar]
  • 89.Rose NR; Woon ECY; Tumber A; Walport LJ; Chowdhury R; Li XS; King ONF; Lejeune C; Ng SS; Krojer T; Chan MC; Rydzik AM; Hopkinson RJ; Che KH; Daniel M; Strain-Damerell C; Gileadi C; Kochan G; Leung IKH; Dunford J; Yeoh KK; Ratcliffe PJ; Burgess-Brown N; von Delft F; Muller S; Marsden B; Brennan PE; McDonough MA; Oppermann U; Klose RJ; Schofield CJ; Kawamura A, Plant growth regulator daminozide is a selective inhibitor of human KDM2/7 histone demethylases. J. Med. Chem 2012, 55 (14), 6639–6643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Luo X; Liu Y; Kubicek S; Myllyharju J; Tumber A; Ng S; Che KH; Podoll J; Heightman TD; Oppermann U; Schreiber SL; Wang X, A selective inhibitor and probe of the cellular functions of Jumonji C domain-containing histone demethylases. J. Am. Chem. Soc 2011, 133 (24), 9451–9456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Xu W; Podoll JD; Dong X; Tumber A; Oppermann U; Wang X, Quantitative analysis of histone demethylase probes using fluorescence polarization. J. Med. Chem 2013, 56 (12), 5198–5202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Wang W; Marholz LJ; Wang X, Novel scaffolds of cell-active histone demethylase inhibitors identified from high-throughput screening. J. Biomol. Screen 2015, 20 (6), 821–827. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Rüger N; Roatsch M; Emmrich T; Franz H; Schüle R; Jung M; Link A, Tetrazolylhydrazides as selective fragment-like inhibitors of the JumonjiC-domain-containing histone demethylase KDM4A. ChemMedChem 2015, 10 (11), 1875–1883. [DOI] [PubMed] [Google Scholar]
  • 94.Małecki PH; Rüger N; Roatsch M; Krylova O; Link A; Jung M; Heinemann U; Weiss MS, Structure-based screening of tetrazolylhydrazide inhibitors versus KDM4 histone demethylases. ChemMedChem 2019, 14 (21), 1828–1839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Johansson C; Velupillai S; Tumber A; Szykowska A; Hookway ES; Nowak RP; Strain-Damerell C; Gileadi C; Philpott M; Burgess-Brown N; Wu N; Kopec J; Nuzzi A; Steuber H; Egner U; Badock V; Munro S; LaThangue NB; Westaway S; Brown J; Athanasou N; Prinjha R; Brennan PE; Oppermann U, Structural analysis of human KDM5B guides histone demethylase inhibitor development. Nat. Chem. Biol 2016, 12 (7), 539–545. [DOI] [PubMed] [Google Scholar]
  • 96.Tschank G; Brocks DG; Engelbart K; Mohr J; Baader E; Günzler V; Hanauske-Abel HM, Inhibition of prolyl hydroxylation and procollagen processing in chick-embryo calvaria by a derivative of pyridine-2,4-dicarboxylate. Characterization of the diethyl ester as a proinhibitor. Biochem. J 1991, 275 (2), 469–476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Joberty G; Boesche M; Brown JA; Eberhard D; Garton NS; Humphreys PG; Mathieson T; Muelbaier M; Ramsden NG; Reader V; Rueger A; Sheppard RJ; Westaway SM; Bantscheff M; Lee K; Wilson DM; Prinjha RK; Drewes G, Interrogating the druggability of the 2-oxoglutarate-dependent dioxygenase target class by chemical proteomics. ACS Chem. Biol 2016, 11 (7), 2002–2010. [DOI] [PubMed] [Google Scholar]
  • 98.Wigle TJ; Swinger KK; Campbell JE; Scholle MD; Sherrill J; Admirand EA; Boriack-Sjodin PA; Kuntz KW; Chesworth R; Moyer MP; Scott MP; Copeland RA, A high-throughput mass spectrometry assay coupled with redox activity testing reduces artifacts and false positives in lysine demethylase screening. J. Biomol. Screen 2015, 20 (6), 810–820. [DOI] [PubMed] [Google Scholar]
  • 99.Korczynska M; Le DD; Younger N; Gregori-Puigjané E; Tumber A; Krojer T; Velupillai S; Gileadi C; Nowak RP; Iwasa E; Pollock SB; Ortiz Torres I; Oppermann U; Shoichet BK; Fujimori DG, Docking and linking of fragments to discover Jumonji histone demethylase inhibitors. J. Med. Chem 2016, 59 (4), 1580–1598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Walters ZS; Aladowicz E; Villarejo-Balcells B; Nugent G; Selfe JL; Eve P; Blagg J; Rossanese O; Shipley J, Role for the Histone Demethylase KDM4B in Rhabdomyosarcoma via CDK6 and CCNA2: Compensation by KDM4A and Apoptotic Response of Targeting Both KDM4B and KDM4A. Cancers (Basel) 2021, 13 (7), 1734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Müller MR; Burmeister A; Skowron MA; Stephan A; Bremmer F; Wakileh GA; Petzsch P; Köhrer K; Albers P; Nettersheim D, Therapeutical interference with the epigenetic landscape of germ cell tumors: a comparative drug study and new mechanistical insights. Clin. Epigenetics 2022, 14 (1), 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Parrish JK; McCann TS; Sechler M; Sobral LM; Hua Ren W; Jones KL; Choon Tan A; Jedlicka P, The Jumonji-domain histone demethylase inhibitor JIB-04 deregulates oncogenic programs and increases DNA damage in Ewing Sarcoma, resulting in impaired cell proliferation and survival, and reduced tumor growth. Oncotarget 2018, 9 (69). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Bayo J; Fiore EJ; Dominguez LM; Real A; Malvicini M; Rizzo M; Atorrasagasti C; García MG; Argemi J; Martinez ED; Mazzolini GD, A comprehensive study of epigenetic alterations in hepatocellular carcinoma identifies potential therapeutic targets. J. Hepatol 2019, 71 (1), 78–90. [DOI] [PubMed] [Google Scholar]
  • 104.Fang Z; Wang T.q.; Li H; Zhang G; Wu X.-a.; Yang L; Peng Y.-l.; Zou J; Li L.-l.; Xiang R; Yang S.-y., Discovery of pyrazolo[1,5-a]pyrimidine-3-carbonitrile derivatives as a new class of histone lysine demethylase 4D (KDM4D) inhibitors. Bioorg. Med. Chem. Lett 2017, 27 (14), 3201–3204. [DOI] [PubMed] [Google Scholar]
  • 105.Wang T; Liu Y; Zhang H; Fang Z; Zhang R; Zhang W; Fan Y; Xiang R, Crystal structures of two inhibitors in complex with histone lysine demethylase 4D (KDM4D) provide new insights for rational drug design. Biochem. Biophys. Res. Commun 2021, 554, 71–75. [DOI] [PubMed] [Google Scholar]
  • 106.Warshakoon NC; Wu S; Boyer A; Kawamoto R; Sheville J; Renock S; Xu K; Pokross M; Zhou S; Winter C; Walter R; Mekel M; Evdokimov AG, Structure-based design, synthesis, and SAR evaluation of a new series of 8-hydroxyquinolines as HIF-1α prolyl hydroxylase inhibitors. Bioorg. Med. Chem. Lett 2006, 16 (21), 5517–5522. [DOI] [PubMed] [Google Scholar]
  • 107.King ONF; Li XS; Sakurai M; Kawamura A; Rose NR; Ng SS; Quinn AM; Rai G; Mott BT; Beswick P; Klose RJ; Oppermann U; Jadhav A; Heightman TD; Maloney DJ; Schofield CJ; Simeonov A, Quantitative high-throughput screening identifies 8-hydroxyquinolines as cell-active histone demethylase inhibitors. PLoS One 2010, 5 (11), e15535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Rai G; Kawamura A; Tumber A; Liang Y; Vogel JL; Arbuckle JH; Rose NR; Dexheimer TS; Foley TL; King ON; Quinn A; Mott BT; Schofield CJ; Oppermann U; Jadhav A; Simeonov A; Kristie TM; Maloney DJ, Discovery of ML324, a JMJD2 demethylase inhibitor with demonstrated antiviral activity. In Probe Reports from the NIH Molecular Libraries Program, National Center for Biotechnology Information (US): Bethesda (MD), 2010. [PubMed] [Google Scholar]
  • 109.Duan L; Rai G; Roggero C; Zhang Q-J; Wei Q; Ma Shi H.; Zhou Y; Santoyo J; Martinez Elisabeth D.; Xiao G; Raj Ganesh V.; Jadhav A; Simeonov A; Maloney David J.; Rizo J; Hsieh J-T; Liu Z-P, KDM4/JMJD2 histone demethylase inhibitors block prostate tumor growth by suppressing the expression of AR and BMYB-regulated genes. Chem. Biol 2015, 22 (9), 1185–1196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Jin C; Yang L; Xie M; Lin C; Merkurjev D; Yang JC; Tanasa B; Oh S; Zhang J; Ohgi KA; Zhou H; Li W; Evans CP; Ding S; Rosenfeld MG, Chem-seq permits identification of genomic targets of drugs against androgen receptor regulation selected by functional phenotypic screens. Proc. Natl. Acad. Sci. U. S. A 2014, 111 (25), 9235–9240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Carter DM; Specker E; Przygodda J; Neuenschwander M; von Kries JP; Heinemann U; Nazaré M; Gohlke U, Identification of a novel benzimidazole pyrazolone scaffold that inhibits KDM4 lysine demethylases and reduces proliferation of prostate cancer cells. SLAS Discov. 2017, 22 (7), 801–812. [DOI] [PubMed] [Google Scholar]
  • 112.Roatsch M; Hoffmann I; Abboud MI; Hancock RL; Tarhonskaya H; Hsu K-F; Wilkins SE; Yeh T-L; Lippl K; Serrer K; Moneke I; Ahrens TD; Robaa D; Wenzler S; Barthes NPF; Franz H; Sippl W; Lassmann S; Diederichs S; Schleicher E; Schofield CJ; Kawamura A; Schüle R; Jung M, The clinically used iron chelator deferasirox is an inhibitor of epigenetic JumonjiC domain-containing histone demethylases. ACS Chem. Biol 2019, 14 (8), 1737–1750. [DOI] [PubMed] [Google Scholar]
  • 113.Fang Z; Liu Y; Zhang R; Chen Q; Wang T; Yang W; Fan Y; Yu C; Xiang R; Yang S, Discovery of a potent and selective inhibitor of histone lysine demethylase KDM4D. Eur. J. Med. Chem 2021, 223, 113662. [DOI] [PubMed] [Google Scholar]
  • 114.Sekirnik R; Rose NR; Thalhammer A; Seden PT; Mecinović J; Schofield CJ, Inhibition of the histone lysine demethylase JMJD2A by ejection of structural Zn(ii). Chem. Commun 2009, (42), 6376–6378. [DOI] [PubMed] [Google Scholar]
  • 115.Kim Y-J; Lee HD; Choi Y-S; Jeong J-H; Kwon HS, Benzo[b]tellurophenes as a potential histone H3 lysine 9 demethylase (KDM4) inhibitor. Int. J. Mol. Sci 2019, 20 (23). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Lohse B; Nielsen AL; Kristensen JBL; Helgstrand C; Cloos PAC; Olsen L; Gajhede M; Clausen RP; Kristensen JL, Targeting histone lysine demethylases by truncating the histone 3 tail to obtain selective substrate-based inhibitors. Angew. Chem. Int. Ed 2011, 50 (39), 9100–9103. [DOI] [PubMed] [Google Scholar]
  • 117.Simon DN; Ulrike L; Magnus B; Silvia AB; Kanchan D; Kamilla M; Daniella I; Jack M; Brian L; Jesper LK; Rasmus PC, Synthesis and characterisation of substrate-based peptides as inhibitors of histone demethylase KDM4C. Protein Pept. Lett 2016, 23 (9), 772–776. [DOI] [PubMed] [Google Scholar]
  • 118.Woon ECY; Tumber A; Kawamura A; Hillringhaus L; Ge W; Rose NR; Ma JHY; Chan MC; Walport LJ; Che KH; Ng SS; Marsden BD; Oppermann U; McDonough MA; Schofield CJ, Linking of 2-oxoglutarate and substrate binding sites enables potent and highly selective inhibition of JmjC histone demethylases. Angew. Chem. Int. Ed 2012, 51 (7), 1631–1634. [DOI] [PubMed] [Google Scholar]
  • 119.Passioura T; Bhushan B; Tumber A; Kawamura A; Suga H, Structure-activity studies of a macrocyclic peptide inhibitor of histone lysine demethylase 4A. Biorg. Med. Chem 2018, 26 (6), 1225–1231. [DOI] [PubMed] [Google Scholar]
  • 120.Kim S-H; Kwon SH; Park S-H; Lee JK; Bang H-S; Nam S-J; Kwon HC; Shin J; Oh D-C, Tripartin, a histone demethylase inhibitor from a bacterium associated with a dung beetle larva. Org. Lett 2013, 15 (8), 1834–1837. [DOI] [PubMed] [Google Scholar]
  • 121.Guillade L; Sarno F; Tarhonskaya H; Nebbioso A; Alvarez S; Kawamura A; Schofield CJ; Altucci L; de Lera ÁR, Synthesis and biological evaluation of tripartin, a putative KDM4 natural product inhibitor, and 1-dichloromethylinden-1-ol analogues. ChemMedChem 2018, 13 (18), 1949–1956. [DOI] [PubMed] [Google Scholar]
  • 122.DeBoer C; Meulman PA; Wnuk RJ; Peterson DH, Geldanamycin, a new antibiotic. J. Antibiot. (Tokyo) 1970, 23 (9), 442–447. [DOI] [PubMed] [Google Scholar]
  • 123.Whitesell L; Mimnaugh EG; De Costa B; Myers CE; Neckers LM, Inhibition of heat shock protein HSP90-pp60v-src heteroprotein complex formation by benzoquinone ansamycins: essential role for stress proteins in oncogenic transformation. Proc. Natl. Acad. Sci. U. S. A 1994, 91 (18), 8324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Wu Q; Lin W; Li Z-M; Rankovic Z; White SW; Chen T; Yang J, A protocol for high-throughput screening of histone lysine demethylase 4 inhibitors using TR-FRET assay. STAR Protocols 2021, 2 (3), 100702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Singh S; Abu-Zaid A; Lin W; Low J; Abdolvahabi A; Jin H; Wu Q; Cooke B; Fang J; Bowling J; Vaithiyalingam S; Currier D; Yun M-K; Fernando DM; Maier J; Tillman H; Bulsara P; Lu Z; Das S; Shelat A; Li Z; Young B; Lee R; Rankovic Z; Murphy AJ; White SW; Davidoff AM; Chen T; Yang J, 17-DMAG dually inhibits Hsp90 and histone lysine demethylases in alveolar rhabdomyosarcoma. iScience 2021, 24 (1), 101996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.van Veen AG; Mertens WK, Toxoflavin, the yellow poison of bongkrek. Recl. Trav. Chim. Pays-Bas Belg 1934, 53, 398–404. [Google Scholar]
  • 127.Li X; Li Y; Wang R; Wang Q; Lu L, Toxoflavin produced by Burkholderia gladioli from Lycoris aurea is a new broad-spectrum fungicide. Appl. Environ. Microbiol 2019, 85 (9), e00106–00119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Daves GD; Robins RK; Cheng CC, Antibiotics. I. Synthesis of 1,6-dimethyl-5,7-dioxo-1,5,6,7-tetrahydropyrimido [5,4-e] as-triazine (Toxoflavin) and related compounds. J. Am. Chem. Soc 1962, 84 (9), 1724–1729. [Google Scholar]
  • 129.Franci G; Sarno F; Nebbioso A; Altucci L, Identification and characterization of PKF118–310 as a KDM4A inhibitor. Epigenetics 2017, 12 (3), 198–205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Letfus V; Jelić D; Bokulić A; Petrinić Grba A; Koštrun S, Rational design, synthesis and biological profiling of new KDM4C inhibitors. Biorg. Med. Chem 2020, 28 (1), 115128. [DOI] [PubMed] [Google Scholar]
  • 131.Souto JA; Sarno F; Nebbioso A; Papulino C; Álvarez R; Lombino J; Perricone U; Padova A; Altucci L; de Lera ÁR, A new family of Jumonji C domain-containing KDM inhibitors inspired by natural product purpurogallin. Front. Chem 2020, 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Rose NR; McDonough MA; King ONF; Kawamura A; Schofield CJ, Inhibition of 2-oxoglutarate dependent oxygenases. Chem. Soc. Rev 2011, 40 (8), 4364–4397. [DOI] [PubMed] [Google Scholar]
  • 133.Sakurai M; Rose NR; Schultz L; Quinn AM; Jadhav A; Ng SS; Oppermann U; Schofield CJ; Simeonov A, A miniaturized screen for inhibitors of Jumonji histone demethylases. Mol. Biosyst 2010, 6 (2), 357–364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Pack LR; Yamamoto KR; Fujimori DG, Opposing chromatin signals direct and regulate the activity of lysine demethylase 4C (KDM4C). J. Biol. Chem 2016, 291 (12), 6060–6070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Upadhyay AK; Judge RA; Li L; Pithawalla R; Simanis J; Bodelle PM; Marin VL; Henry RF; Petros AM; Sun C, Targeting lysine specific demethylase 4A (KDM4A) tandem TUDOR domain – A fragment based approach. Bioorg. Med. Chem. Lett 2018, 28 (10), 1708–1713. [DOI] [PubMed] [Google Scholar]
  • 136.Wagner EK; Nath N; Flemming R; Feltenberger JB; Denu JM, Identification and characterization of small molecule inhibitors of a plant homeodomain finger. Biochemistry 2012, 51 (41), 8293–8306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Zhang Y; Yang H; Guo X; Rong N; Song Y; Xu Y; Lan W; Zhang X; Liu M; Xu Y; Cao C, The PHD1 finger of KDM5B recognizes unmodified H3K4 during the demethylation of histone H3K4me2/3 by KDM5B. Protein Cell 2014, 5 (11), 837–850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Bhushan B; Erdmann A; Zhang Y; Belle R; Johannson C; Oppermann U; Hopkinson RJ; Schofield CJ; Kawamura A, Investigations on small molecule inhibitors targeting the histone H3K4 tri-methyllysine binding PHD-finger of JmjC histone demethylases. Biorg. Med. Chem 2018, 26 (11), 2984–2991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Tumber A; Nuzzi A; Hookway ES; Hatch SB; Velupillai S; Johansson C; Kawamura A; Savitsky P; Yapp C; Szykowska A; Wu N; Bountra C; Strain-Damerell C; Burgess-Brown NA; Ruda GF; Fedorov O; Munro S; England KS; Nowak RP; Schofield CJ; La Thangue NB; Pawlyn C; Davies F; Morgan G; Athanasou N; Müller S; Oppermann U; Brennan PE, Potent and selective KDM5 inhibitor stops cellular demethylation of H3K4me3 at transcription start sites and proliferation of MM1S myeloma cells. Cell Chem. Biol 2017, 24 (3), 371–380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Vinogradova M; Gehling VS; Gustafson A; Arora S; Tindell CA; Wilson C; Williamson KE; Guler GD; Gangurde P; Manieri W; Busby J; Flynn EM; Lan F; Kim H.-j.; Odate S; Cochran AG; Liu Y; Wongchenko M; Yang Y; Cheung TK; Maile TM; Lau T; Costa M; Hegde GV; Jackson E; Pitti R; Arnott D; Bailey C; Bellon S; Cummings RT; Albrecht BK; Harmange J-C; Kiefer JR; Trojer P; Classon M, An inhibitor of KDM5 demethylases reduces survival of drug-tolerant cancer cells. Nat. Chem. Biol 2016, 12 (7), 531–538. [DOI] [PubMed] [Google Scholar]
  • 141.Kruidenier L; Chung C.-w.; Cheng Z; Liddle J; Che K; Joberty G; Bantscheff M; Bountra C; Bridges A; Diallo H; Eberhard D; Hutchinson S; Jones E; Katso R; Leveridge M; Mander PK; Mosley J; Ramirez-Molina C; Rowland P; Schofield CJ; Sheppard RJ; Smith JE; Swales C; Tanner R; Thomas P; Tumber A; Drewes G; Oppermann U; Patel DJ; Lee K; Wilson DM, A selective jumonji H3K27 demethylase inhibitor modulates the proinflammatory macrophage response. Nature 2012, 488 (7411), 404–408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Vazquez-Rodriguez S; Wright M; Rogers CM; Cribbs AP; Velupillai S; Philpott M; Lee H; Dunford JE; Huber KVM; Robers MB; Vasta JD; Thezenas M-L; Bonham S; Kessler B; Bennett J; Fedorov O; Raynaud F; Donovan A; Blagg J; Bavetsias V; Oppermann U; Bountra C; Kawamura A; Brennan PE, Design, synthesis and characterization of covalent KDM5 inhibitors. Angew. Chem. Int. Ed 2019, 58 (2), 515–519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Mares A; Miah AH; Smith IED; Rackham M; Thawani AR; Cryan J; Haile PA; Votta BJ; Beal AM; Capriotti C; Reilly MA; Fisher DT; Zinn N; Bantscheff M; MacDonald TT; Vossenkamper A; Dace P; Churcher I; Benowitz AB; Watt G; Denyer J; Scott-Stevens P; Harling JD, Extended pharmacodynamic responses observed upon PROTAC-mediated degradation of RIPK2. Commun. Biol 2020, 3 (1), 140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Iida T; Itoh Y; Takahashi Y; Yamashita Y; Kurohara T; Miyake Y; Oba M; Suzuki T, Design, synthesis, and biological evaluation of lysine demethylase 5 C degraders. ChemMedChem 2021, 16 (10), 1609–1618. [DOI] [PubMed] [Google Scholar]
  • 145.Arkin Michelle R.; Tang Y; Wells James A., Small-molecule inhibitors of protein-protein interactions: progressing toward the reality. Chem. Biol 2014, 21 (9), 1102–1114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Bushweller JH, Targeting transcription factors in cancer — from undruggable to reality. Nat. Rev. Cancer 2019, 19 (11), 611–624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Singh B; Sarli VN; Lucci A, Sensitization of resistant breast cancer cells with a Jumonji family histone demethylase inhibitor. Cancers (Basel) 2022, 14 (11). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Mar BG; Chu SH; Kahn JD; Krivtsov AV; Koche R; Castellano CA; Kotlier JL; Zon RL; McConkey ME; Chabon J; Chappell R; Grauman PV; Hsieh JJ; Armstrong SA; Ebert BL, SETD2 alterations impair DNA damage recognition and lead to resistance to chemotherapy in leukemia. Blood 2017, 130 (24), 2631–2641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Banelli B; Daga A; Forlani A; Allemanni G; Marubbi D; Pistillo MP; Profumo A; Romani M, Small molecules targeting histone demethylase genes (KDMs) inhibit growth of temozolomide-resistant glioblastoma cells. Oncotarget 2017, 8 (21), 34896–34910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Bayo J; Tran TA; Wang L; Pena-Llopis S; Das AK; Martinez ED, Jumonji inhibitors overcome radioresistance in cancer through changes in H3K4 methylation at double-strand breaks. Cell Rep. 2018, 25 (4), 1040–1050 e1045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Yao B; Yang Q; Yang Y; Li Y; Peng H; Wu S; Wang L; Zhang S; Huang M; Wang E; Xiong P; Luo T; Li L; Jia S; Deng Y; Deng Y, Screening for active compounds targeting human natural killer cell activation identifying Daphnetin as an enhancer for IFN-gamma production and direct cytotoxicity. Front. Immunol 2021, 12, 680611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Son J; Huang S; Zeng Q; Bricker TL; Case JB; Zhou J; Zang R; Liu Z; Chang X; Darling TL; Xu J; Harastani HH; Chen L; Gomez Castro MF; Zhao Y; Kohio HP; Hou G; Fan B; Niu B; Guo R; Rothlauf PW; Bailey AL; Wang X; Shi PY; Martinez ED; Brody SL; Whelan SPJ; Diamond MS; Boon ACM; Li B; Ding S, JIB-04 has broad-spectrum antiviral activity and inhibits SARS-CoV-2 replication and coronavirus pathogenesis. mBio 2022, e0337721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Tran TA; Zhang QJ; Wang L; Gonzales C; Girard L; May H; Gillette T; Liu ZP; Martinez ED, Inhibition of Jumonji demethylases reprograms severe dilated cardiomyopathy and prolongs survival. J. Biol. Chem 2022, 298 (2), 101515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Zhang QJ; Tran TAT; Wang M; Ranek MJ; Kokkonen-Simon KM; Gao J; Luo X; Tan W; Kyrychenko V; Liao L; Xu J; Hill JA; Olson EN; Kass DA; Martinez ED; Liu ZP, Histone lysine dimethyl-demethylase KDM3A controls pathological cardiac hypertrophy and fibrosis. Nat Commun 2018, 9 (1), 5230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Cercek A; Lumish M; Sinopoli J; Weiss J; Shia J; Lamendola-Essel M; El Dika IH; Segal N; Shcherba M; Sugarman R; Stadler Z; Yaeger R; Smith JJ; Rousseau B; Argiles G; Patel M; Desai A; Saltz LB; Widmar M; Iyer K; Zhang J; Gianino N; Crane C; Romesser PB; Pappou EP; Paty P; Garcia-Aguilar J; Gonen M; Gollub M; Weiser MR; Schalper KA; Diaz LA, PD-1 blockade in mismatch repair-deficient, locally advanced rectal cancer. N. Engl. J. Med 2022, 386 (25), 2363–2376. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES