Introduction

As one of the major causes of mortality worldwide, cancer challenges global public health. According to cancer statistics 2018, one in every five men and one in every six women would develop cancer during their lifetime.1 Moreover, there were approximately 19.3 million new cancer cases and 10.0 million cancer deaths in 2020.2 The occurrence of cancer entails a series of genetic mutations that favor uncontrollable tumor growth. It is believed that various factors collectively contribute to cancer, and there is no one single explanation for tumorigenesis. For instance, cancers can be caused by internal factors such as spontaneous DNA mutations or external environmental factors. Epigenetic alterations are a class of features common to cancer progression, reversibly modulating oncogenesis through chromatin compaction, and are susceptible to external or internal environmental factors.3 The term “epigenetics” was originally introduced by Dr. Waddington to describe the hereditary alterations in cell phenotypes that were independent of DNA sequence change.4 Following decades of research, the definition of epigenetics has reached a consensus that epigenetics is the chromatin-based event that modulates DNA-templated processes.5

Composed of DNA and the surrounding nucleosomes, chromatin is a constantly-changing structure that responds to external environments. Each nucleosome contains an octamer of four histones (H2A, H2B, H3, and H4), the post-translational modifications (PTMs), which influences chromatin compaction and subsequently modulates the transcription levels of different genes.6 The histone modifications at specific residues include acetylation, methylation, phosphorylation, citrullination, ubiquitination, ADP-ribosylation, deamidation, formylation, O-GlcNAcylation, propionylation, butyrylation, crotonylation, and proline isomerization, controlling gene expression during the development of diseases.7 It is thus not surprising that molecules regulating the deposition and removal of histone modifications are actively involved in oncogenesis and inflammation response.8,9

The methylation of lysine residues on histone proteins represents a class of PTMs. Lysine residues of histones can be either mono-, di-, or tri-methylated (Kme1, Kme2, and Kme3, respectively) by enzymes that recognize methyl marks on histone proteins. The different methylation status of histones leads to the recruitment of binding proteins with varying affinities.10,11 The methylation process of lysines is accomplished by the histone lysine methyltransferases (KMTs), also referred to as “epigenetic writers”, whereas the removal of methyl groups on lysine relies on lysine demethylases (KDMs), referred to as “erasers”.12,13 KDMs are classified into two families according to their action mechanism, the flavin adenine dinucleotide (FAD)-dependent amine oxidases and the Jumonji C (JmjC) domain-containing (JMJD) demethylases. Figure 1 presents the phylogenetic tree of histone demethylase members of JMJD family proteins.

Fig. 1
figure 1

Phylogenetic tree of histone demethylase members of JMJD family proteins. Figure was created with Biorender (www.bioender.com)

The human JMJD protein family consists of more than 30 members, most of which have been identified with histone lysine demethylase activity. The JMJD family of KDM enzymes function as Fe2+ and 2-oxoglutarate-dependent dioxygenases and are able to demethylate histone lysines at different methylated states (Kme1, Kme2, and Kme3).14,15 The signature structure of JMJD proteins is a ~170 amino acids long Jumonji C (JmjC) domain where their complexation with Fe2+ occurs.15,16 The activity of JMJD proteins requires the involvement of cofactors oxygen and 2-oxoglutarate (or α-ketoglutarate), resulting in the sensitivity of JMJD activity to the metabolic changes within cells.17 Two consecutive chemical reactions are implicated in the lysine demethylation process, including the hydroxylation of the methylated ε-amino group and formaldehyde release. In addition to methylated lysine residues, JMJD proteins also demonstrate their hydroxylation activities on amino acid residues of aspartate, asparagine, histidine, arginine, and unmethylated lysine, and tRNA.18

The first protein identified with JmjC domain-based catalytic activity was HIF1AN (hypoxia-inducible factor 1 subunit alpha inhibitor), a hydroxylase of asparagine residues.19,20 This has led to speculation that JMJD proteins could hydroxylate methylated lysine residues and thereby exhibit their demethylation ability.21 Soon thereafter, a number of JMJD proteins were detected for their histone lysine demethylase activities, which collectively form a large heterogenous JMJD protein family.14,22 The classification of JMJD family members can be based on their molecular weight (>100 kDa or <100 kDa), the specificity of lysine demethylation, or the existence of functional domains. Considerable attention has converged on those JMJD proteins reported with histone lysine demethylase activity.23 Though some other JMJD family members are not catalytically active, such as JARID2, which contains amino acid mutations critical for cofactor binding, they are still essential for the multiple biological processes.24 In this review, we will focus on the role of JMJD family proteins in cancer and inflammation, including the intensively studied histone lysine demethylases and the understudied group of JMJD members. The representative diagram of the demethylating activities of JMJD family members in cancer and inflammation is presented in Fig. 2.

Fig. 2
figure 2

Representative diagram of the demethylating activities of JMJD family members in cancer and inflammation: JMJD3 demethylates tri-methylated Lys 27 on histone H3 and thus affects the transcription of inflammation-associated gene. JMJD6 is a 2OG oxygenase and catalyze methylarginine demethylation of histone H3/H4 residues, which regulates the transcription of cancer-related pathway genes such as MAPK signaling. Figure was created with Biorender (www.bioender.com)

JMJD proteins in cancer

Growing evidence has demonstrated the aberrant expression of JMJD proteins in cancer and inflammatory diseases, which might serve as an underlying mechanism for the initiation and progression of such diseases. The inhibiting and promoting effects of JMJD family proteins in different cancer types are summarized in Table 1 and Fig. 3.

Table 1 The inhibiting and promoting activities of JMJD family proteins in cancer
Fig. 3
figure 3

The inhibiting and promoting effects of JMJD family proteins in different cancer types. Figure was created with Biorender (www.bioender.com)

JMJD1

The JMJD1 subfamily includes JMJD1A, JMJD1B, and JMJD1C (or KDM3A, B, and C, respectively), which all contain 2 conserved domains required for catalytic activities,25,26 JmjC catalytic center, and a C6 zinc finger. JMJD1A is 59.64% identical to JMJD1B in terms of amino acids. However, both JMJD1A and JMJD1B share less than 50% of their amino acid sequences with JMJD1C, suggesting that JMJD1C is evolutionarily distinct.18 Though these JMJD1 proteins have been reported to demethylate mono- and dimethylated lysine 9 on histone H3 (H3K9), the trimethylated H3K9 is not a substrate for JMJD1 proteins.27 It was recently suggested that the demethylation activity of JMJD1 might include more histone lysines in addition to H3K9.28,29 JMJD1 proteins have long been found to regulate normal homeostasis, and here we mainly focus on their roles in oncogenesis.

On one hand, JMJD1A functions as a tumor-suppressive factor, with Jmjd1a knockout resulting in increased microvessel formation and the expression alterations of angiogenesis-related genes, such as the downregulation of anti-angiogenic factors.30 JMJD1A also suppresses the proliferation of gastric cancer cells by regulating its target gene runt-related transcription factor 3 (RUNX3).31 On the other hand, the overexpression of JMJD1A was found to correlate with increased metastasis and unfavorable prognosis of colorectal cancer (CRC) and gastric cancer.32,33 JMJD1A may promote CRC progression via Wnt/β-catenin signaling where it coactivates downstream targets of β-catenin32,34,35 or coactivating STAT3 transcription factor.36 JMJD1A expression is significantly elevated in bladder carcinomas compared with adjacent noncancerous tissues, which promotes the G1/S cycle transition by regulating HOXA1 gene transcription.37 Another mechanism through which JMJD1A promotes urinary bladder cancer progression is the increased glycolysis induced by JMJD1 through coactivation of HIF1α.38

In prostate cancer, JMJD1A promotes the proliferation and survival of prostate cancer cells via the regulation of c-Myc expression at both transcriptional and post-translational levels.39 In other words, JMJD1A not only enhances c-Myc transcriptional activity but also prevents the degradation of c-Myc protein.39 JMJD1A was found to promote the expression of factors that mediate DNA repair and radioresistance of prostate cancer cells, making JMJD1A a potential therapeutic target to improve the response of prostate cancer cells to chemotherapies, radiotherapies, and PARP inhibitors.40 JMJD1A promotes the formation of alternative splicing of AR variant 7 (AR-V7), a key mechanism by which prostate cancer cells develop resistance to androgen deprivation therapy.41

JMJD1B is located in the 5q31 chromosomal locus, the deletion of which is often seen in myelodysplasia and acute myeloid leukemia (AML). Lower expression of JMJD1B in AML patients indicated worse prognosis,42,43 and JMJD1B was thus viewed as a tumor suppressor for AML.44 The underlying mechanism may be the JMJD1B-facilitated degradation of PML/RARα, a critical event in the pathogenesis of acute promyelocytic leukemia (APL).43 JMJD1B also appears as a tumor suppressor in CRC, the histone methylation of which is regulated by PRL-3, an essential metastasis gene of CRC.45

JMJD1C was initially identified in undifferentiated spermatogonia, the knockout of which resulted in increased apoptosis of germ cells in mice.46 In the field of cancer, JMJD1C functions as an oncogenic factor for AML by promoting cell survival and self-renewal. For example, intracranial germ cell cancer was characterized by germline missense mutations in JMJD1C.47 JMJD1C is overexpressed in colon cancer tissues and increases colon cancer metastasis via the inactivation of the ATF2 pathway.48 On the contrary, JMJD1C functions as a tumor suppressor in esophageal cancer, which downregulates cancer cell proliferation by targeting YAP1 gene expression via H3K9me2 demethylation.49 Interestingly, in JMJD1C-knockout mice, the global H3K9 methylation level remained unchanged, and it thus is postulated whether JMJD1C displays H3K9 demethylase activity.50

JMJD2/KDM4

JMJD2 is one of the largest JMJD subfamilies and is comprised of JMJD2A-D proteins, also referred to as KDM4A-D. JMJD2 family members take active parts in multiple physiological processes, including cell proliferation, migration,51 gene transcription,52 and genome stability.53 In embryonic stem cells (ESCs), KDM4B and KDM4C interact with pluripotency factors such as Sox2, Oct4, c-Myc, and Klf4, thereby regulating cell proliferation and stem-cell features.54,55 It was recently reported that JMJD2A, B, and C were all crucial to the survival of acute myeloid leukemia cells,56 but the individual role of each member varied significantly.

JMJD2A uses trimethylated H3K9 and H3K36 as demethylating substrates. Interestingly, demethylating efficiency of JMJD2A on H3K9me3 is 5-fold higher than on H3K36me3 and higher on trimethylated than dimethylated H3K9/H3K36.57 The dual role of JMJD2A in transcription, with both stimulating and repressing effects on gene transcription, has drawn considerable research attention. JMJD2A may directly bind to transcription factors58 or interact with nuclear receptor corepressor to suppress gene transcription.59,60 The oncogenic effect of JMJD2A was first observed in breast cancer, with approximately 60% of breast tumors identified with JMJD2A overexpression.61,62,63 JMJD2A is also important for androgen and estrogen receptor (ER) activities based on its catalytic activity.64 The absence of JMJD2A in ER-positive or ER-negative breast cancer cells decreased the expression of ER target genes such as the c-Jun and cyclin D1, leading to aberrant cell proliferation.65 A recent meta-analysis revealed the differential expression of JMJD2 in breast cancer, with JMJD2A/D overexpression predominantly observed in basal-like breast cancer and JMD2B in ER-positive luminal-type breast cancer.66

JMJD2A also plays a functional role in a number of other cancers. In lung cancer, JMJD2A decreased the transcription of tumor suppressor gene CHD5 to block cellular senescence, which ultimately stimulated cellular transformation.67 Likewise, the upregulation of JMJD2A expression was later reported in prostate cancer and bladder cancer tissues.68 In bladder cancer, JMJD2A promoted epithelial-mesenchymal transition (EMT) by modulating SLUG expression. However, contradictory results were reported that lower JMJD2A intensity was observed in bladder cancer tissue samples, predicting significantly worse overall survival.69 JMJD2A promoted the growth and protein synthesis of gliomas via phosphoinositide-dependent kinase-1 (PDK1)-mediated Akt-mTOR pathway activation.70 Notably, there appeared to be no difference between JMJD2A upregulation or downregulation in the growth of cervical carcinoma.71 These results suggested that JMJD2A might preferentially stimulate the growth of specific tumor types.

JMJD2B and JMJD2C share similar structures and action specificity to JMJD2A.72 It remains unclear whether the catalytic activity of JMJD4B is lower than other JMJD2 members because the different sizes of recombinant JMJD4B proteins would affect the measurement results.73 It is widely accepted that JMJD2B supports the carcinogenesis of ER-positive tumors because it is, in fact, an ER target gene.74,75,76 Though JMJD2B/C is upregulated in breast cancers at mRNA levels, higher JMJD2B expression was observed in ER-positive than ER-negative breast cancer, whereas the reverse applied for JMJD2C.77,78 As a downstream target of the pluripotency factor Oct4, JMJD2C promoted not only the proliferation of ER-negative breast cancer cells but also cancer-stem-cell features such as mammosphere formation in non-transformed breast cancer cell lines,79 suggesting the important role of JMJD2C in the maintenance of cancer stem cells. JMJD2B is associated with invasion and metastasis of gastric cancer by inducing EMT.80 JMJD2B and β-catenin collectively promote the transcription of β-catenin target gene vimentin via H3K9 demethylation.80 The overexpression of JMJD2B was found to correlate with the abundance of p-c-Jun in gastric cancer, which is predictive of poor survival.81 In classical Hodgkin lymphoma, the elevated expression of JMJD2B and JMJD2D was also associated with aggressive subtypes and suboptimal treatment response to radiation.82

In CRC patients, the overexpression of JMJD2B in CRC specimens indicates a poor prognosis. It has been demonstrated that JMJD2B accelerated CRC progression based on its interaction with TRAF6, which leads to TRAF6-mediated AKT activation.83 A more recent study identified a novel epigenetic mechanism for the progression of CRC, where JMJD2B enhances the transcription of small GTPase TC10-like (TCL), leading to a malignant phenotype of CRC cells.84 Additionally, under glucose deficient conditions, JMJD2B sustained the autophagy-derived amino acids in CRC cells via the epigenetic regulation of LC3B, thereby promoting the aggressiveness of CRC.85 Similar to its action mechanism in gastric cancer, JMJD2B supports the gene transcription induced by β-catenin, thereby contributing to the tumorigenesis of CRCs.86 The invasion of CRC cells may also be attributed to the immune escape via the KDM4B/HOXC4/PD-L1 axis.87

A unique member of the JMJD2 family, JMJD2D, is only half the size of JMJD2A-C due to its lack of PHD and Tudor domains.88 Compared with JMJD2A-C, which uses H3K36 as demethylating substrates, JMJD2D has a different substrate-binding specificity and acts on dimethylated H1.4K26 rather than trimethylated H1.4K26.89 JMJD2D also demethylates H3K9me2 and H3K9me3 but less efficiently demethylates H3K9me1.90,91 The expression of JMJD2D in the margins of pancreatic tumors was indicative of earlier recurrence in patients.92 JMJD2D is highly expressed in liver cancer, and its demethylase-independent inhibition on p53 tumor suppressor promotes liver cancer initiation and progression.93 The chemical inhibition of JMJD2 by ML324 enhances cell apoptosis of hepatocellular carcinoma via the unfolded protein response and Bim upregulation.93

Compared with noncancerous colon tissues, JMJD2D is highly expressed in CRC tissues and promotes tumor growth and invasion of CRC. The crosstalk between JMJD2D and β-catenin activates the transcription of β-catenin target genes in CRC cells.94 The colon tumorigenesis by JMJD2D can be mediated by Hedgehog signaling. In addition, a recent study suggested that JMJD2D enhanced CRC progression by activating HIF1 signaling and subsequent cell glycolysis.95 The activation of the HIF1 signaling pathway by JMJD2D can be based on three mechanisms: (1) JMJD2D upregulates mTOR expression, thus promoting HIF1α translation; (2) JMJD2D upregulates HIF1β transcription; (3) JMJD2D interacts with HIF1α to induce glycolytic gene expression.95 JMJD2D is also involved in the immune escape of CRC cells by upregulating PD-L1 expression, providing a new strategy to improve response to anti-PD-1/PD-L1 immunotherapies.96 Bioinformatical analyses revealed that for two potential JMJD2 members, the gene products of JMJD2E and JMJD2F were similar to those of JMJD2D, but JMJD2E and JMJD2F are more likely to be pseudogenes.97

Importantly, the role of JMJD2/KDM4 proteins in tumorigenesis is especially addressed in colorectal cancer. JMJD2A promoted cell growth of colon cancer by increasing cell proliferation and at the same inhibiting apoptosis.58 JMJD2B is also overexpressed in CRC tissues, the inhibition of which promotes cell apoptosis providing a potential therapeutic strategy.98 JMJD2B could be induced under a hypoxic environment in a HIF-1α-dependent manner in CRC cells. Under such circumstances, the expression of several hypoxia-inducible genes was upregulated by JMJD2B via demethylation of H3K9me on their promoters.99 It was recently suggested that Wnt-mediated CRC metastasis is partially dependent on JMJD2 to form an epigenetic complex that activates disintegrin and metalloproteinase (ADAM) transcription.100

JMJD3/KDM6B

The JMJD3 gene is located on chromosome 17p13.1101,102 and 88% homologous to the gene histone demethylase gene UTX (ubiquitously repeat transcribed tetratricopeptide repeat on the X chromosome).103 UTX was the first identified mutated histone demethylase gene associated with cancer104 and can remove the methyl groups from di-trimethylated H3K27.103,105,106,107 The location of JMJD3 is adjacent to p53 as well, a tumor suppressor, the mutation of which is a frequent event in cancer. It was found that JMJD3 might also directly interact with p53.108,109 The role of JMJD3 on cancer is highly controversial, with tumor inhibitory effects on CRC, hepatic cancer, pancreatic cancer, glioma and B-cell lymphoma, and tumor-promoting effect on cancers such as renal breast, prostate, and ovarian cancer.

The regulating effect of JMJD3 in cancer is partially attributed to its role in the EMT of cancer cells. Transforming growth factor-β (TGF-β) is a well-characterized multipotent cytokine that inhibits tumor cell proliferation at the early stage but induces EMT at the late stage of cancer progression.110 The increased expression of JMJD3 was proved to be associated with metastatic capacities of ovarian cancer via the increased expression of TGF-β.111 Moreover, JMJD3 promoted EMT of Ras-mutated lung cancer cells via TGF-β-mediated Smad stimulation.112 In line with the in vitro studies, non-small cell lung cancer (NSCLC) patients with high JMJD3 expression in tumor specimens displayed higher risks of lymphatic and distant metastasis and poor overall survival.113 Further studies provided a potential mechanism through which JMJD3-mediated EMT of cancer cells. JMJD3 enhanced TGF-β-induced EMT by upregulating the EMT-related gene, SNAI1, in invasive breast cancer.114

Previous evidence also investigated the important yet poorly defined role of JMJD3 in the metastasis of CRC, the second most lethal cancer worldwide in 2020.2 The aberrant expression of Wnt/β-catenin pathway molecules is often observed in a wide spectrum of cancers and is believed to be associated with epigenetic modulation of key gene promoters in colon cancer.115 JMJD3 has a dual role in CRC as a tumor suppressor and tumor activator. JMJD3 is under-expressed in CRC patient specimens, and the absence of the JMJD3 in the tumor is characterized as a marker of poor clinical outcome in CRC.116 Earlier evidence identified JMJD3 as a downstream target of vitamin D metabolite 1α,25-dihydroxyvitamin D(3) (1,25(OH)(2)D(3)) in colon cancer cells which mediates the effects of 1,25(OH) 2D 3 on a subset of EMT-inducer genes such as ZEB1, ZE, B2, and SNAI1.117 The expression of JMJD3 was found to inversely correlate with that of SNAI1 in colon cancer tissues, and the inhibition of JMJD3 abolished 1,25(OH)(2)D(3)-induced β-catenin transcriptional activity.118 On the other hand, JMJD3 promotes the expression of the epithelial cell adhesion molecule (EpCAM) gene in CRC based on its histone promotor demethylation function.119 The aberrant activation of NOTCH1, and the subsequent increase in Ephrin type-B receptor 4 (EPHB4) expression, are considered a hallmark of CRC progression. The intracellular NOTCH domain led JMJD3 to the EPHB4 enhancer region, and modified the chromatin architecture by regulating the H3K27me3 level, which ultimately resulted in EPHB4 activation.120 Recently, JMJD3 has been reported to control tumor immunosuppression. JMJD3-mediated H3K27me3 reduced the production of Th1-type chemokines CXCL9 and CXCL10, mediators of effector T-cell trafficking in colon cancer.121

In accordance with the tumor-supportive action of JMJD3 in CRC, the regulation of JMJD3 on cancer development based on its demethylation activities can also be found in a series of other cancer types. The brain is the most investigated organ for JMJD3 regulation. Patients with pediatric brainstem glioma often experience a decrease in H3K27me3, and the methylation maintenance by JMJD3 inhibition thus becomes an important treatment strategy.122,123 JMJD3 expression was elevated in glioma relative to normal tissues, and the inhibiting JMJD3 demethylation reduced tumor cell proliferation and migration, and at the same time enhanced apoptosis.124 In contrast, some reports addressed JMJD3 as a tumor suppressor in glioma via modulating the expression of the key transcription factors, including the p53.125 As discussed earlier, JMJD3 may directly interact with p53 and regulate its activity independently of chromatin modification, leading to glioblastoma stem-cell (GSC) differentiation.108 Moreover, JMJD3 expression is partially determined by STAT3 (signal transducer and activator of transcription 3), which binds to and inhibits the promotor region of JMJD3. Once JMJD3 expression was resumed from STAT3 inhibition, JMJD3 reduced the formation of neurosphere and cell proliferation of GSC.126 One of the treatment strategies for glioblastoma is an aptamer which targets the ligand of platelet-derived growth factor receptor A (PDGFRα), leading to decreased STAT3 and increased JMJD3 expression, and subsequent upregulation of p53.127 Collectively, these results suggested the central position of JMJD3 in the STAT3-JMJD3-p53 signal network that regulates glioma progression.128

The occurrence of prostate cancer (PC) is highly relevant to histone modifications such as methylation,129 and the fact that an H3K27 methyltransferase has been found to indicate prostate cancer progression further supports the oncogenic role of histone methylation status at H3K27 in prostate cancer.130 PC is hormone-dependent cancer with overexpression of androgen receptor (AR).131 A wide breadth of literature has suggested the link between JMJD3, H3K27me3, and the AR metabolic pathway.132,133,134,135,136 Increased transcriptional level of JMJD3 was reported in metastatic prostate cancer,105 with higher expression in AR-positive compared to AR-negative PC cell lines.132 The signature genes activated by JMJD3 in PC include (O-methylguanine-DNA methyltransferase (MGMT), transformer 2 alpha homolog (TRA2A), and 2 small nuclear RNA auxiliary factor 1 (U2AF1), and ribosomal protein S6 kinase A2 (RPS6KA2)), identified as signature genes in PC.137

The overexpression of JMJD3 is often observed in germinal center B (GC-B) cells in Hodgkin lymphoma (HL)138 and diffuse large B-cell lymphoma (DLBCL).139,140 Following JMJD3 blockade treatment, the H3K27me3 level on target genes was significantly decreased in HL, supporting the conclusion that JMJD3 is involved in the development of HL.138 In DLBCL, JMJD3 promotes the phosphorylation of proteins mediating the B-cell receptor (BCR) signaling. It affects its downstream B-cell lymphoma 6 protein (BCL6), facilitating normal B-cell survival and lymphogenesis of B-cell non-Hodgkin lymphoma (NHL).141 Moreover, JMJD3 is involved in the treatment response to chemotherapies, with JMJD3 inhibitors demonstrating significant chemo-sensitization on B cells.139 Recently, a potential link was established between JMJD3 demethylase and cyclin-dependent kinase 9 (CDK9), the abnormal expression of which was a frequent event in DLBCL. The use of CDK9 inhibitors reduced JMJD3 expression, which specifically elevated the trimethylation of H3K27.142

JMJD3 is also involved in a broad spectrum of cancers such as PML/RARα-positive leukemic, where JMJD3 activates expression of the homeobox (HOX) gene via interaction with PML-RARα fusion protein.143 In neuroblastoma which is mainly induced by the activation of oncogenes and the failure of neural crest cell differentiation, blocking JMJD3 may regulate the expression of several key differentiation genes such as the v-myc myelocytomatosis viral-related oncogene (MyCN).144 Based on these intriguing findings, further studies are warranted to clarify the precise role and action mechanisms of JMJD3 in cancer under different circumstances.

JMJD4

JMJD4 is a recently identified histone demethylase homologous to JMJD6. However, compared with JMJD6, which has been intensively studied, experimental work to characterize the role of JMJD4 in cancer has lagged far behind. Currently, the only reported enzymatic activity of JMJD4 is the hydroxylation of the lysine residue of eukaryotic release factor 1 (eRF1).18 Though the inhibition of eRF1hydroxylation led to less efficient transcriptional termination, researchers failed to identify any physiological consequences.145 An initial investigation suggested significantly higher JMJD4 expression in tumor tissues than in normal tissues of the colon and hepatic cancer and the differential expression of JMJD4 protein in colon cancer of different histological grades and metastasis status.146 Recent research revealed the increased expression of JMJD4 expression in renal cancer, which might be a prognostic marker in renal cancer patients.147 Thus, more studies are needed to delineate the role of JMJD4 in cancer.

JMJD5 and JMJD7

JMJD5/ KDM8 shuttles between the cytoplasm and the nucleus148 exhibit a wide range of enzymatic activities, including the demethylation of H3K36me2,149,150 hydroxylation at the C3 of arginine residues,151,152,153 and proteolysis.154,155 Recent research has cast doubt on the demethylation of JMJD5 H3K36me2 based on the crystal structure results that the catalytic center of JMJD5 is not favorable to the accommodation of methylated lysine residues.151,156,157,158 Further, no valid in vivo evidence has been reported for the arginine hydroxylation activities of JMJD5.153 In keeping with JMJD5, JMJD7 is able to hydroxylate the C3 position of lysine residues and at the same time, cleaves arginine methylated histone as a protease.154 The hydroxylation of DRG1/2, two GTPases involved in ribosome biogenesis,159 by JMJD7 enhanced the binding of DRG1/2 to RNA. Noteworthy, the occupancy of JMJD7 at gene promoter regions negatively regulates osteoclast differentiation, suggesting the critical role of JMJD7 in bone formation and turnover.160 The divalent cation-dependent protease activities of JMJD5 and JMJD7 preferentially cleave the tails of H2, H3, and H4 bearing methylated arginine. Like other aminopeptidases, JMJD5 and JMJD7 digest the C-terminal products following the initial specific cleavage, providing a new repertoire for removing histone tails with methylated arginine residue.154 Interestingly, histones such as H3 and H4 and their arginine methylated isoforms were increased in cells lacking either JMJD5 or JMJD7 in vivo.154

JMJD5 is highly expressed in breast cancer cell lines, the knockdown of which resulted in tumor cell growth arrest.149 It was recently suggested that there is significantly lower mRNA expression of JMJD5 in breast cancer, hepatocellular carcinoma, and lung cancer, but a higher expression in stomach adenocarcinoma than in normal tissues. Accordingly, high JMJD5 expression indicated a good prognosis in breast cancer, hepatocellular carcinoma, and lung cancer but a poor prognosis in stomach adenocarcinoma.161 In this study, JMJD5 expression was also related to the abundance of infiltrating immune cells in tumors, which might jointly serve as a prognostic marker.

In cancer cells, JMJD5 promotes the transcription of PKM2-HIF-1α target genes that mediates glucose metabolism, leading to increased glucose uptake and lactate secretion.162 JMJD5 has been identified as a binding partner for p53 tumor suppressor and positively regulates cell proliferation and cell cycle.163 Thus, in oral squamous cell carcinoma, the downregulation of JMJD5 significantly induces apoptosis and reduces tumor metastasis via p53/NF-κB pathway.164 A recent study described potential mechanisms for the regulation of both androgen-responsive and metabolic genes by JMJD5 in castration-resistance of prostate cancer (CRPC) cells. JMJD5 can either interacts with androgen receptor (AR) and modulates androgen response, or with PKM2 to regulate tumor metabolism under androgen-deprived conditions.165 Moreover, JMJD5 inhibition prevented cancer-stem-cell-like mediated by cancer upregulated gene 2 (CUG2).166 In pancreatic cancer however, JMJD5 negatively regulates c-Myc expression, which suppresses tumor cell proliferation and glycolytic metabolismr.167

Likewise, deletion of JMJD7 also impaired the viability of prostate cancer cells,168 and reduced colony formation of breast cancer cells.154 JMJD7 also promoted cell survival of head and neck squamous cell carcinoma (HNSCC) by modulating the phosphorylation of protein kinase B.169 It was recently hypothesized that the loss of function of mixed-lineage leukemia gene (MLL) 1 fusion, a major cause of pediatric leukemia, coupled with the failed conversion of H3K4me1 to H3K4me3, might trigger the malignant transformation of cells, suggesting a potential role of JMJD5 and JMJD7 in leukemia development.170

JMJD6

JMJD6 is a 47.5 kDa protein with 403 amino acids. JMJD6 is a monomer and can be in the trimeric, pentameric or larger oligomeric form in solution and in fibril form in the absence of its poly-Ser sequence.171 Earlier reports refereed JMJD6 as a surface marker on macrophages, fibroblasts and epithelial cells, originally named PSR (phosphatidylserine receptor).172,173 Later studies found that JMJD6 was in fact predominantly located in the cellular nucleus both in cells endogenously expressing JMJD6 and JMJD6-transfected cells.174 As PSR was further proved as a nuclear 2-oxoglutarate (2OG)-and Fe(II)-dependent oxygenase,175 it was later renamed to JMJD6.176 In mouse models, JMJD6 deficiency resulted in neonatal lethality and serious defects in the development of organs, independent of its apoptotic cell removal activities.177

A key catalytic activity of JMJD6 is the arginine demethylation, both on mono-methylarginine and dimethylarginine residues of histones. To date JMJD6 is the only enzyme reported with a potential arginine demethylation activity in vivo.178 Recent evidence has presented non‐histone targets of arginine demethylation by JMJD6, including RNA helicase A, estrogen receptor α (ERα), tumor necrosis factor receptor‐associated factor 6 (TRAF6), and the transcription factor PAX3 and heat‐shock protein 70 (HSP70).179,180,181,182 However, some studies cast doubt on the function of JMJD6 as a histone arginine demethylase in cells such as endothelial cells.183 Researchers failed to identify arginine methylation at H4R3 in JMJD6-knockdown endothelial cells,184 which was further supported by the crystal structure analysis that JMJD6 structure was not conduction to demethylation activities.185

As a multi-functional enzyme intensively involved in chromosomal rearrangement and gene transcription, JMJD6 functions as arginine demethylase and lysyl hydroxylase,186 and even tyrosine kinase of histones.187 In glioblastoma and neuroblastoma, JMJD6 forms protein complexes with N-Myc and BRD4 (Bromodomain-containing protein 4), which is important for gene transcription of a number of genes including E2F2, N-Myc and c-Myc.188,189 As a tumorigenesis factor for neuroblastoma, JMJD6 is highly expressed in human neuroblastoma tissues and the knockdown of JMJD6 decreased neuroblastoma cell proliferation and tumor progression in vivo, suggesting the potential of JMJD6 as a therapeutic target in neuroblastoma.190 Moreover, JMJD6 is a critical regulator of AR splice variant 7 (AR-V7) which mediates the endocrine resistance in advanced prostate cancer.191

Accumulating evidence suggested that increased JMJD6 expression in breast cancer cells was associated with increased tumor growth and metastasis.192,193,194 According to analyses from patient tumor samples, the expression level of JMJD6 varies among breast cancer subtypes. For instance, ER‐positive tumors exhibited significantly lower JMJD6 expression than ER‐negative tumors, which explained the fact that JMJD6 was consistently related to ER‐negative diseases.193 However, in this study, no significant correlation between the JMJD6 level and the prognosis was identified.193,194 Furthermore, Claudin‐low breast tumors displayed the highest JMJD6 expression, followed by basal-like, HER2‐enriched and luminal B subtypes, with the lowest expression detected in luminal A subtype.193

The oncogenic role of JMJD6 in oral squamous cell carcinoma is potentially attributed to stem‐like properties mediated by JMJD6,195,196,197 which is assumed as a key factor for cancer recurrence and treatment failure.198 JMJD6 is crucial to melanoma progression as well, the mutation, amplification, or deletion of which indicates unfavorable prognosis.199 Interestingly, a study identified a novel post-translational modification of P53 by JMJD6 independent of its histone arginine demethylation activity, where JMJD6 antagonized p53 acetylation and repressed its following transcriptional activity in colon cancer.200 Other cancer types that have been reported to be affected by JMJD6 expression levels include lung cancer,201,202 hepatic cancer,203 and ovarian cancer,204 where high level of JMJD6 expression correlates with increased cell proliferation, invasiveness, and poor clinical outcomes.

JMJD8

JMJD8 is evolutionarily distant from the other members of the JMJD family,18 which contains a JmjC domain at 74–269 amino acid residues with no other recognizable protein domains.205 JMJD8 is mainly localized at the endoplasmic reticulum and reportedly involved in angiogenesis and cell metabolism.206,207 Previous research has demonstrated that JMJD8 JMJD8 functions as a positive regulator of TNF-induced NF-κB signaling.205 The knockdown of JMJD8 upregulated AKT/NF-κB/COX-2 pathway and enhanced Ku70/Ku80 expression in cancer cells, thereby regulating cell proliferation and their responses to cancer treatments that induced DNA damage.208 A recent study investigated the prognostic value of 8 glycolysis-related genes in HNSCC and identified JMJD8 as a protective gene for HNSCC.209 Another study verified that JMJD8 functioned as an oncogene in CRC which promoted cell proliferation and EMT through the NF-κB pathway.210 Likewise, JMJD8 promoted carcinogenesis of NSCLC cells by maintaining EGFR stability and the downstream PI3K/AKT signaling pathway,211 which accorded with a recent finding that JMJD8 could modulate tumor EMT via AKT activation.212 Thus, JMJD8 is a potential prognostic marker and therapeutic target for cancer patients. However, JMJD8 has not been thoroughly studied and its precise role in cancer remains to be elucidated.

JMJD10/MDIG

Given the different identification sources and multiple biological functions, JMJD10 is also frequently referred to as RIOX2, Mina53/Mina, NO52 or MDIG (mineral dust-induced gene).213,214,215,216 JMJD10 was initially detected in alveolar macrophages of coal miners and its expression can be induced by environmental cancer risk factors such as silica, smoke and arsenic.217 The inverse correlation between MDIG expression and H3K9me3 level in lung tumors218 supported the role of MDIG as a histone demethylase and an epigenetic regulator in a number of cancer types.219,220,221 One structural study failed to prove the histone demethylase activity of MDIG, but rather identified its hydroxylase activity toward the ribosomal protein L27a (RPL27a).222 For the first time, the recent study for the first time identified MDIG as an antagonist for histone methylation repressors, suggesting the potential of MDIG as a new target for cancer therapy.223

Many cancers have been identified with overexpression of MDIG relative to normal tissues, such as breast cancer,224 colon cancer,225,226 lymphoma.220,227,228,229,230,231 The screening results of the expression pattern of lung cancer revealed that 90 percent of lung cancers displayed elevated MDIG expression level.230 An underlying mechanism for MDIG-induced invasion and metastasis of lung cancer cells may be the destabilization of β-catenin and subsequent suppression of EMT-related genes.232 MDIG was frequently overexpressed in hepatocellular carcinoma (HCC) which was associated with higher histological grades, potentially modulating HCC progression via MDIG/H3K9me3/p21 pathway.233 In some cancers, higher MDIG expression is predictive of worse prognosis.234 Interestingly, the prognostic value of MDIG may vary in the same cancer types. For instance, increased MDIG expression is associated with longer OS in breast cancer patients with lymph node or distal metastasis.224 Likewise, overexpression of MDIG is indicative of prognosis only in patients at stages I/II, but not in stages III/IV patients.216 These results suggest that MDIG may be an oncogene that promotes tumor growth at early stages and a tumor-suppressive gene that reduces the metastatic capacity of tumor at late stages.

JARID1(Jumonji And AT-Rich Interaction Domain Containing 1)/ KDM5

Before identifying of their histone demethylase activities, JARID1 family members were initially reported to play important roles in stem-cell biology and congenital disease. Four JARID1 members that are upregulated in cancers have been identified so far, JARID1A, JARID1B, JARID1C and JARID1D. JARID1A was first identified by screening a library of cDNA that interacted with retinoblastoma gene product (pRb),235 and its crosstalk with pRb reinforced the transcription repression on cell differentiation by retinoblastoma family.236

JARIDA is highly expressed in pediatric acute megakaryoblastic leukemia (AMKL), and its C-terminal PHD finger forms a fusion gene with NUP98.237,238 This fusion impairs myeloblast differentiation and promotes self-renewal of progenitor cells, which is important to leukemogenic transformation.239 The expression of JARIDA is also upregulated in breast cancer,240,241,242,243,244 prostate cancer,245 and gastric cancer,246,247,248,249 potentially by enhancing tumor cell proliferation and metastasis. In triple-negative breast cancer (TNBC), the anti-tumoral effect of blocking JARIDA impaired cell cycle progression and p16/p27-mediated senescence,244 which accorded with previous results on gastric cancer where JARID1A repressed cyclin-dependent kinase inhibitors including p16, p21, and p27.249 The oncogenic effect of JARIDA on PC is partially attributed to its demethylation activities on H3K4, which decreases the expression of KLF4 and E-cadherin, and facilitates cell proliferation and metastasis.250 JARIDA can also induce tumor metastasis of TNBC independent of its demethylase-dependent function, by promoting integrin β−1 (ITGB1) expression.240 Furthermore, JARIDA is responsible for the generation of therapeutic responses in cancer.251 One such example is breast cancer resistance to trastuzumab and erlotinib induced by JARIDA.252 On the contrary, JARIDA improved the treatment response of melanoma cells to immune checkpoint blockade,253 suggesting that the role of JARIDA in treatment response may vary based on tumor types and drug types.

JARID1B was initially regarded as an oncogene in breast cancer,254 the overexpression of which was significantly associated with poor prognosis,255 despite later research identifying its suppressive activities on the invasion of TNBS cells.256 JARID1B is upregulated in a wide range of cancers including prostate,257 hepatocellular,258 head and neck cancers259 and ovarian cancer,260except for melanomas with relatively low JARID1B expression.261 JARID1B expression has been reported to confer stem cell-like features to cancer cells,262 and regulate oxidative metabolism.263 JARID1B may also reduce the progression by suppressing the genome-wide H3K4me3 hyper-methylation in leukemias.264 Recent evidence suggested that JARID1-targeted inhibitors could overcome cisplatin resistance to platinum-based chemotherapeutics in melanoma.265 Likewise, CPI-455, the first tool compound selectively targeting the JARID1 family, inhibited the stem cell-like properties of oral cancer.266

JARID1C is an X-linked gene,267 the aberrant function of which leads to X-linked retardation.268 It has been well established that JARID1C plays a dual role both as a tumor promoter and a tumor suppressor. For instance, in clear cell renal cell carcinoma (CCRCC), mutations in JARID1C gene lead to its functional loss and the preferential occurrence of CCRCC in males.269,270 JARID1C could impair the development of papilloma virus-related malignancies by forming a complex with viral E2 that suppressed the E6 and E7 viral oncoprotein promoters.271

Earlier studies referred to JARID1D as a minor histocompatibility antigen on the Y chromosome.272 The downregulation, mutation, or loss of JARID1D was recently shown in metastatic PC and CCRCC.273In hormone-sensitive PC, the interaction between JARID1D and androgen receptor inhibits the transcriptional activation of AR target genes and loss of JARID1D may result in treatment failure due to dysregulating AR signaling.274 These efforts may only represent the tip of the iceberg regarding the role of JARID1D in cancer progression, and more studies are warranted to address the biological contributions of JARID1 to cancer.

JARID2 (Jumonji And AT-Rich Interaction Domain Containing 2)

JARID2 is often described as and probably the most widely studied a PRC2-associated factor.275,276,277,278,279 PRC2 is a protein complex consisting of 4 core subunits, including the enhancer of zeste homolog 1 or 2 (EZH1/2), embryonic ectoderm development (EED), suppressor of zeste 12 (SUZ12) and retinoblastoma associated protein 46/48, (RbAP46/48), also known as RBBP4/7.280 PRC2 is responsible for the mono-, di and tri-methylation of H3K27, with approximately 70% of H3 histones being methylated by PRC2.281,282,283,284

Though JARID2 is a founder member of the JMJD protein family,285 it lacks the essential residues required for enzymatic activity, making its JmjC domain inactive.286,287 Since its discovery, the role of JARID2 in mammalian development has mainly converged in the embryonic stem cell pluripotency. Accumulating evidence has reported its function in embryonic lethality, based on different genetic backgrounds of the mutant strain.288 Apart from the JmjC domain, JARID2 contains a JmjN domain and two other domains with DNA-binding capacity which is likely to be independent of its crosstalk with PRC2.278

Except for the its function in the embryological context,289,290 JARID2 was also dysregulated in cancer and considered as an oncogene that promotes cancer progression. Through inhibiting the overactivation of AKT induced by phosphatase and tension homolog (PTEN), JARID2 facilitated EMT and invasion of HCC cells.291 It is thus not surprising that the knockdown of JARID2 reduced the TGF-β-mediated EMT in colon and lung cancer cells.292 A recent study demonstrated the essential role of the LINC021/IMP2/JARID2 signaling axis in CRC tumorigenesis where LINC021 enhanced the mRNA stability of JARID2.293 In bladder cancer, JARID2 promoted the proliferation, migration, invasion and sphere-forming capacities of bladder cancer cells.294 Previous studies suggested that tumor cells undergoing EMT are more likely to be resistant to cisplatin.295,296 Researchers later found that JARID2 was involved in developing cisplatin resistance in non-small cell lung cancer via upregulation of Notch1.297 Nevertheless, JARID2 is not always an oncogene that facilitates tumorigenesis. JARID2 can also function as a hematopoietic tumor suppressor that limits the self-renewal of multipotent progenitor cells and prevents the transformation of nonmalignant blood disorders such as myeloproliferative neoplasms and myelodysplastic syndromes, into AML.298

UTX/KDM6A and UTY

KDM6A or UTX was first identified in 2007 together with JMJD3 and ubiquitously transcribed tetratricopeptide repeat on chromosome Y (UTY) as a group of H3K27 demethylases.299,300 Whereas UTX is an X-linked protein with demethylating activities on H3K27me2/3, UTY is the Y-linked homolog of UTX which shares similar structures with UTX with minimal demethylase activity due to a mutation in the JmjC catalytic domain.107,301

UTX is one of a few cancer suppressors that escape X inactivation, leading to a predominant occurrence rate in the male population.270 According to the analyses of 4,742 human tumor specimens.302 UTX is highly mutated across various cancers, including acute lymphoblastic leukemia,303 chronic myelomonocytic leukemia,304 bladder cancer,305 medulloblastoma,306 prostate cancer,307 and renal carcinoma.308 The proliferation of cancer cells was reduced when inactivating UTX mutations were resumed with the addition of wild-type UTX. An increased mutational rate of UTX was observed from 10.7% to 21.6% in pancreatic cancer samples. The knockdown or inactive mutations of UTX increased TP63 expression, which was considered a key driver of pancreatic cancer.309

On the contrary, specimen analysis of patients with oral tongue squamous cell carcinoma (OTSCC) suggested that UTX expression in tumor tissues may predict poor survival outcomes in theses patients.310 In NSCLC cells, UTX is regarded as an oncogene which promotes cell proliferation and migration, and its expression is modulated by the EGFR-STAT3 axis.311 The promoting effect of UTX on lung cancer oncogenesis is mainly mediated through the upregulation of EZH2, and the UTX-deficient lung cancer is preferentially sensitive to EZH2 inhibitors.312 In addition, immunohistochemistry staining results revealed that UTX was highly expressed in CRC tissues and promotes CRC cell proliferation and maintains G0/G1 cell cycle progression via upregulating KIF 14 and pAKT.313 UTX positively regulates E-cadherin expression via modulating H3K27 demethylation and acetylation, activating the transcription of the E-cadherin at its promoter regions.314

In breast cancer, depletion of UTX resulted in upregulation of Myc-dependent expression of EMT factors, including SNAI and ZEB1/2.315 Thus, by forming a transcriptional repressive complex with LSD1, HDAC1 and DNMT1, UTX is a tumor suppressor and a negative regulator of EMT-induced CSC-like properties in breast cancer.314 However, pro-tumor functions of UTX were observed in the ER + subtype of breast cancer cells where the transactivation of UTX and estrogen receptor (ER) forms a feed-forward loop in response to hormone treatments.316

UTY on the other hand, is less frequently mutated in cancer than UTX317 and had less tumor-suppression effect than UTX.309,318,319 It was recently reported that UTY displayed weaker tumor-suppression abilities than UTX in leukemia, which was further reduced by the deleting the UTY cIDR (residues 498–795).320 Similar to UTX, the depletion of UTY promoted cell proliferation of urothelial bladder cancer cells321 and 12% of urothelial bladder carcinomas were identified with the absence of UTY.322 Moreover, the knockout of both UTX and UTY had a synergistic effect on the increase of proliferation, which might be attributed to the loss of dosage-dependent suppression effect of UTX/UTY in urothelial cancer.

JMJD proteins in inflammation

Histone modifications lead to significant alterations in genome structures and functions. Figure 4 presents the signaling pathways involved in the regulation of cancer and inflammation by JMJD family members and their crosstalks.

Fig. 4
figure 4

The signaling pathways involved in the regulation of cancer and inflammation by JMJD family members and their crosstalks. Figure was created with Biorender (www.bioender.com)

JMJD1 and JMJD2

Under hypoxic conditions, the increased HIF-1α expression promotes the inflammatory injury of endothelial cells, which is independent on the NF-κB pathway. In JMJD1A-knockdown human umbilical vein endothelial cells (HUVECs), a number of genes involved in inflammation and the oxidative stress pathways were significantly downregulated.323

The involvement of JMJD2 in inflammation is best represented by JMJD2D which mediates inflammatory responses elicited by cytokines such as tumor necrosis factor α (TNFα), and consequently reshapes the immune microenvironment. TNFα is a pro-inflammatory cytokine produced by monocytes during acute inflammation and is implicated in a range of events leading to cell necrosis or apoptosis.324 TNFα was able to induce JMJD2D expression in dendritic cells and macrophages,23 and the demethylation of H3K9 by JMJD2D in turn participated in the TNFα response.325 In response to colon injury caused by inflammatory bowel disease (IBD), TNF-α secreted by macrophages activates the NF-κB signaling, upregulating JMJD2D in the colon epithelial cells.326

The expression of JMJD2B is significantly upregulated in gastric epithelial cells during H. pylori infection via β-catenin signaling.327 β-catenin directly binds to the promoter region of JMJD2B gene and activates its transcription. The upregulated JMJD2B, together with NF-κB, binds to COX-2 promoter to stimulate its transcription via demethylation of H3K9me3.327 Vascular inflammation is regarded as an preliminary step towards multiple human diseases, and its contributing factors remain incompletely defined. A recent study investigated the epigenetic changes during the transformation of vascular smooth muscle cells (VSMCs) into osteoblast-like cells in response to inflammation, and found that JMJD2B was the downstream target of IL-6/STAT3 axis, suggesting the pathogenic role of JMJD2B during chronic inflammation.328 In a LPS-induced vascular inflammation model, JMJD2A promotes the transactivation of pro-inflammatory cytokines, facilitating the binding of SET1A to NF-κB promoter.329

JMJD3

As discussed earlier, JMJD3 together with UTX and UTY, belongs to the KDM6 family. However, as UTX and UTY are mainly involved in developmental processes, JMJD3 is implicated in regulating of inflammation and cellular senescence.330,331,332 JMJD3 is usually expressed at a low levels under normal conditions, but its expression is drastically increased by inflammatory stresses such as hypoxia inducers and oncogenic factors.333,334,335,336

JMJD3 interacts with distinct transcription factors and potently promotes the expression of inflammatory genes through H3K27me2,3 demethylation.337,338 One such example is the JMJD3-induced activation of NF-κB signaling genes.339,340,341 The knockdown of JMJD3 in human monocytic cells altered the expression profile of inflammatory genes in including chemokines, CD40 signaling and NF-κB-related inflammatory genes. In particular, JMJD3 knockdown and the subsequent enrichment of H3K27me3 at the promoter regions of NF-κB signaling genes suppresses the transcription of genes such as such as monocyte chemoattractant protein-1 (MCP-1) and IL-1β.340 In glomerular mesangial cells, the activation of the NF-κB/JMJD3 signaling pathway could promote high glucose-induced inflammation.342

JMJD3 also participates in the transcription process independent of its demethylase activity.343,344 For instance, STAT1 and STAT3 stimulated the transcription, which subsequently enhanced the expression of Lipopolysaccharide (LPS)-induced inflammatory genes.345 Besides, JMJD3 is also implicated in the SMAD3-mediated TGF-β signaling pathway.346

JMJD3 is involved in several inflammation-related diseases such as rheumatoid arthritis (RA)347,348 where JMJD3 modulates the inflammation persistence and angiogenesis of RA via transcription factor GATA4.349 Given that the H3K27 level is elevated in the midbrain of aged mice, JMJD3 might reshape the immune microenvironment of Parkinson’s disease.350 In Parkinson’s disease, JMJD3 enhances the M1 pro-inflammatory response by suppressing the anti-inflammatory microglia M2 phenotype, resulting in increased neuronal cell death.350,351 In diabetic peripheral tissues, JMJD3 mediates the chronic activation of macrophages, providing another rationale for using histone demethylase inhibitors for the treatment of nonhealing diabetic wounds.352

JMJD3 also plays a pivotal role in cell response to bacteria, parasites, or virus infection. JMJD3 modulates the recovery of murine macrophages from exposure to the lethal anthrax toxin.353 During the latency stage of herpes simplex virus 1 (HSV-1) infection, JMJD3 prevents the reactivation of HSV-1 in sensory neurons by decreasing H3K27me3.354 JMJD3 deficiency in CD4+ T cells leads to the accumulation of T cells in the thymus, and reduced T-cell trafficking to the secondary lymphoid organs. The underlying mechanism for the regulation is the binding of JMJD3 to the Pdlim4 promoter which modulates its expression to affect T cell trafficking.355

JMJD6

JMJD6 is most highly expressed in innate immune cells. One of its target genes by its arginine demethylase activity is tumor necrosis factor receptor-associated factor 6 (TRAF6), which can be both methylated and demethylated at different arginine sites.181 TRAF6 could be involved in the pathogenesis of a variety of autoimmune diseases,356 and the reversible arginine methylation status of TRAF6 by JMJD6 thus provides a novel mechanism for regulation of innate immune pathways. A recent study reported a potential underlying mechanism for the pathogenesis of neuropathic pain. The overexpression of JMJD6 suppressed the activation of NF-κB signaling peripheral nerve injury, suggesting its therapeutic value in neuropathic pain.357 JMJD6 is also involved in viral RNA replication. Immunoprecipitation assays confirmed a physical interaction between recombinant JMJD6 and DHX9,179 which is required to replicate the foot-and-mouth disease virus (FMDV) in cells.358

The impact of JMJD6 on transcriptional regulator Aire reflects its critical role in the spontaneous development of multi-organ autoimmunity in mice, such as thymus and T cell development.359 For example, in patients with chronic hepatitis B virus infection, T lymphocytes usually experience a decrease in JMJD6 expression.360 It is thus postulated that the “exhausted” T cells when exposed to chronic inflammation can be partially attributed to aberrant JMJD6 expression. In fact, the deficiency of JMJD6 in normal peripheral blood mononuclear cells specifically inhibited CD4+ T cell proliferation and is associated with an increased level of cyclin-dependent kinase inhibitor 3 (CDKN3),360 a suppressor of cell cycle progression.361 However, it remains incompletely defined whether the activity of JMJD6 in T cell exhaustion is cell autonomous.362

JMJD8

A recent gene expression profiling analysis demonstrated the highly enriched expression of JMJD8 in adipocytes, which is affected by metabolic and nutritional status.363 JMJD8 expression in turn exerts its regulatory effect on the expression of a series of pro-inflammatory genes, thereby triggering inflammation responses. Importantly, functional interaction between JMJD8 and IRF3, a pro-inflammatory factor involved in adipocyte inflammation and insulin sensitivity, suggested that JMJD8 might be a junction bridging adipocyte insulin sensitivity and inflammation.363 Moreover, JMJD8 functions as a positive regulator of TNF-induced NF-κB signaling,205 which regulates a large array of genes involved in multiple immune and inflammatory responses.364

MDIG

Despite conflicting results regarding the impact of MDIG on Th2 development, it has been well established that MDIG is involved in Th2 response-related atopic asthma and parasitic helminth infection. A genetic case-control study suggested that the T allele of MDIG correlated with an increased risk of atopic asthma, a disease typically driven by pulmonary inflammation.365,366 MDIG deficiency extenuates airway hyper-responsiveness and pulmonary inflammation, possibly by controlling IL-4 production.367 MDIG may also promote silica-induced lung fibrosis by altering the balance between Th17 and Treg cells.368 More recently, MDIG mediates the response to environmental exposure to COVID-19, making it a therapeutic target of COVID-19 ameliorates the pulmonary symptoms.369 Further studies are warranted to elucidate the underlying mechanisms for the involvement of MDIG in pulmonary inflammation.

JARID2

The role of JARID2 in inflammation is best characterized by its function in Crohn’s disease (CD), the most common type of inflammatory bowel disease. An intricate series of pathological factors are associated with the onset of CD, but the exact molecular mechanisms remain incompletely defined.370,371 Regulatory B cells producing IL-10 facilitate intestinal homeostasis, which potently inhibits mucosal inflammatory responses of intestines.372 Patients with Crohn’s disease have a decreased level of regulatory B cells373,374 and the deficiency in B10 cells is reportedly related to CD development.375 A recent study demonstrated the increased IL-10 production by B cells mediated by JARID2 which promotes H3K27me3 binding to the IL10 promoter regions. This provides a novel molecular explanation for the pathogenesis of B10 cells in CD patients.376

Another study introduced a novel mechanism through which inflammatory cytokine interferon-γ (IFN-γ) and class II transactivator (CIITA) collectively reset the fate of post-inflammation muscle cells.377 IFN-γ has been found to prevent muscle development during inflammation.378 Circulating IFN-γ increased PCR recruitment in a JARID2-dependent manner, thereby suppressing muscle-specific genes. Moreover, as one of the target genes of miR-155, JARID could attenuate theTh2 and Th17-mediated airway inflammation.379

UTX

As described earlier, it remains to be elucidated whether UTX expression contributes to the female predominance of autoimmune diseases. Previous analyses investigated UTX expression in CD4+ T lymphocytes of female versus male mice,380,381 and suggested that the X escape of UTX is involved in a wide range of immune response genes, providing a potential explanation for the female susceptibility to autoimmune disease.382 The role of UTX in innate immune responses described so far relies on its H3K27me2/3 demethylation activities in macrophages, which promotes pro-inflammatory cytokine transcription such as IL-6 and IFN-β.383 NF-κB signaling can be activated by UTX, leading to increased secretion of macrophage migration inhibitory factor in neural stem cells.384 Thus, UTX could be recognized as a protecting marker that improves neurological function recovery after spinal cord injury.

Besides, UTX is necessary for the differentiation of CD4+ T cells to Tfh cells during chronic virus infection. The depletion of UTX in mice promoted H3K27 methylation level, decreased the gene expression at Tfh-related genetic loci, and led to deficient virus-specific IgG production.385 UTX gene mutations are often associated with bladder cancer. Recently, the absence of UTX was reported to induce activation of inflammatory pathways that contributes to bladder cancer in cooperation with p53 dysfunction.386 The loss of UTX in CD4+ T cells also aggravates allergic contact dermatitis in mice.387

JMJD proteins as therapeutic targets

Given that the JMJD class of histone demethylase is involved in various physiological and pathological processes, specific JMJD inhibitor would be an attractive strategy, the utility of which should not be limited to combating cancer but also the treatment of inflammatory disorders such as asthma. However, due to its highly polar 2-OG binding pocket, the development of small-molecule inhibitors for the JMJD family has lagged behind, with several JMJD inhibitors being reported but functionally inactive.388,389 We summarized known inhibitors targeting JMJD family proteins evaluated for the treatment of cancer and inflammatory disease (Table 2).

Table 2 Inhibitors targeting JMJD family proteins evaluated for the treatment of cancer and inflammatory disease

JMJD2 inhibitor

An early study described a variety of inhibitor scaffolds with the capacity to suppress 2-OG-dependent JMJD2 histone demethylases, which would facilitate the establishment of small-molecule probes for the identification of enzyme functions in epigenetic signaling.390 Known JMJD2 inhibitors can be classified as either 2-OG cofactor mimics, substrate-competitors, metal cofactor inhibitors, and peptide inhibitors.391 Cofactor mimics competitively binding to Fe(II) at the catalytic site of JMJD2 proteins and modifies the availability of the 2-OG cofactor required for cancer cell metabolism. This class of JMJD2 inhibitors includes fumarate and succinate, which have long been identified as 2-OG antagonists.392

As the overexpression of JMJD2 is frequently observed in breast cancers, previous studies mainly analyzed the therapeutic potential of JMJD2 inhibitors in breast cancer. For instance, a JMJD2 inhibitor, NCDM-32B, effectively decreased cell growth of basal breast cancer cell lines.66 With structure-based drug design, a novel JMJD2 inhibitor QC6352 was developed, which potently suppressed the proliferation, sphere formation, and in vivo tumor growth of TNBC, as well as PDX models of colon cancer.389 Moreover, QC6352 abrogated EGFR expression, thereby overcoming therapeutic resistance in breast cancer.393 Recently TACH101, a pan inhibitor of the JMJD2 subfamily was introduced. This compound exhibited high inhibitory efficacy on four KDM4 isoforms (A-D) and was able to induce cell apoptosis of esophageal cancer, TNBC, and CRC cell lines. Animal studies presented a 4.4-fold lower tumor-initiating cell frequency by TACH101.394 In lung cancer, JMJD2 inhibition by either JMJD2 selective inhibitor ML324 or pan-JMJD inhibitor JIB04 could overcome cisplatin resistance, potentially by preventing ATR-Chk1 replication checkpoint.395 Furthermore, the combined treatment of JMJD2 inhibitors and LSD1 inhibitors may represent a more effective strategy for the enhancement of chemotherapy efficacy.396

The 5-chloro-8-hydroxyquinoline (5-c-8HQ), also referenced under CAS 5852-78-8, is a well-studied JMJD2D inhibitor used in multiple researches. The treatment of 5-c-8HQ in mice leads to significantly smaller and fewer colitis-associated tumors.94 JMJD2D inhibition using 5-c-8HQ decreased the self-renewal capacities of liver cancer stem-like cells, thereby suppressing live cancer progression.397 It has also been reported that 5-c-8HQ works in synergy with Hedgehog inhibitor vismodegib to suppress CRC tumorigenesis and cell proliferation.326

A group of tumor-initiating cells (TIC) were isolated from patient samples of esophageal squamous cell carcinoma (ESCC) which is characterized by stem cell-like features. Importantly, JMJD2C expression was upregulated in this subpopulation, suggesting the potential of JMJD2C inhibition in eliminating ESCC TIC compartment.398 JMJD2C was reported to confer stem-cell-like characteristics in ESCC cells and caffeic acid (3,4-dihydroxycinnamic acid, CA) is able to suppress its demethylation activity.399 An ongoing clinical trial aims to investigate the efficacy and safety of caffeic acid for the treatment of esophageal cancer (NCT03070262).400 It is the only clinical trial up to date that is registered on www.clinicaltrials.gov to assess the efficacy of JMJD protein inhibitors for cancer treatment. In this trial, 240 patients with advanced esophageal squamous cell cancer (ESCC) were randomized into two arms: coffeic acid treatment (300 mg, tid, po) or placebo treatment. Patients will be followed every year and the clinical outcomes will be recorded as overall survival and progression-free survival. The application of JMD2 inhibitor is not limited to cancer treatment. ML324 has potent anti-viral activity against both herpes simplex virus (HSV) and human cytomegalovirus (hCMV) infection and recurrence, suggesting the therapeutic value of chromatin-based inhibitors against viral infection.401,402 ML324 was also tested for the treatment of the depression-like condition in mice by increasing the repressive histone methylation in the nucleus accumbens.403

JMJD3 and UTX inhibitor

A research team reported the first selective JMJD3 inhibitor, supporting its role as a therapeutic target in epigenetic drug discovery. With the optimization of a series of compounds obtained from the screening of a compound collection,404 a JMJD3 and UTX specific inhibitor GSK-J1 was developed with a half-maximum inhibitory concentration (IC50) of 60 nM.405,406 GSK-J3 was refined based on GSK-J1 with substitution at the para position to the pyridine nitrogen and improved access to solvent. Later, the acid groups of GSK-J1 and GSK-J2 were concealed with ethyl esters, which derived new compounds GSK-J4 and GSK-J5.

The inhibition of JMJD3 by GSK-J4 increased the level of H3K27 methylation and demonstrated potent antitumor efficacy in glioma and leukemia where the H3K27me dysregulation occurs recurrently.122,407,408 The underlying mechanism for the potent efficacy of GSK-J4 in leukemia might be its downregulation of Cyclic-AMP response element-binding protein.409 Recent evidence suggested a potential combinatory effect of GSK-J4 and decitabine in leukemia cells by inducing cell cycle arrest, cell apoptosis and PKC-α/p-bcl2 pathway inhibition.410 In glioma cells, GSK-J4 is a combination partner for deacetylase inhibitor panobinostat in glioma cells, and for doxorubicin in KRAS-mutant anaplastic thyroid cancer.411,412 GSK-J4 is also involved in the radiosensitization of diffuse intrinsic pontine glioma, an aggressive pediatric brainstem tumor by inducing DNA repair deficiency.408 Other cancers that responded to GSK-J4 included BC,413 PC,131 lung cancer,414 hemangiosarcoma,415 SMARCA4-mutant cancer,416 Ewing sarcoma,417 and chondrosarcoma418 where GSK-J4 displayed potent antitumor effect.

GSK-J4 cannot be perceived solely as an antitumor drug, as it is also applied for the treatment of inflammatory diseases such as inflammatory colitis by suppressing the inflammatory potential and increasing the generation of tolerogenic dendritic cells.419 GSK-J4 selectively reduced intracellular labile iron in dopaminergic neurons, and this neuroprotection is based on its epigenetic mechanism, suggesting the therapeutic potential of GSK-J4 for Parkinson’s disease.420

JMJD6 inhibitor

Given that JMJD6 has been reported with both demethylation and hydroxylation activities, both of which require the presence of Fe (II) and 2-OG and occur at the same active sites, it is thus speculated that the inhibition of JMJD6 can rely on the targeting of either arginine demethylation or lysyl hydroxylation.

To date, three JMJD6 inhibitor candidates have been proposed which all remain at the preclinical stage. The first JMJD6-targeting inhibitor, WL12, was developed following silico protocol by targeting the druggable 2OG-binding site. The inhibition of JMJD6 enzymatic activity by WL12 lead to decreased cell proliferation.421 Another JMJD6 inhibitor, SKLB325, significantly suppressed the proliferation and induced cell apoptosis of ovarian cancer in a dose-dependent manner. The study further suggested the colocalization of JMJD6 with p53 in the nucleus, upregulating p53 and its downstream effectors.204 SKLB325 also sensitizes renal cell carcinoma cells to sunitinib and works synergistically with sunitinib in inhibiting RCC growth.422 Recently, a research team performed molecular docking and retrieved a new JMJD6 inhibiting compound J2, the optimization of which yielded a more potent JMJD6 inhibitor 7p. The IC50 value of 7p against JMJD6 was 0.681 μM, with minimal activity against other JMJD family proteins.423

JARID1 inhibitor

The contributions of JARID1 to cancer progression have derived respective countermeasures targeting JARID1. A series of pan-JARID1 inhibitors were optimized from compound 1, a hit initially designed to specifically target JARID1C. The optimization of compound 1 led to compound 20 which is highly selective for JARID1 enzymes and able to induce a global increase in H3K4me3 level.424 Another study combined a high throughput screening hit with an established scaffold, and developed a novel JARID1 inhibitor 1,7-naphthyridones, which is more selective to JARID1 over JMJD2 related isoforms.425

The first JARID1-specific inhibitor is CPI-455, with 200-fold higher selectivity for JARID1 than for JMJD2.426 The inhibition of JARID1B by CPI-455 reduced the stem cell-like features of oral squamous cell carcinoma cells, but cells also displayed demethylase-independent activities refractory to inhibition.266 JARID1A is highly expressed in drug tolerant persister (DTP) cells, a subpopulation of tumor cells that contributes to the growing number of drug resistant cells427 such as TMZ-resistant glioblastoma cells.428 Given that drug tolerance of tumor cells was partially dependent on demethylase activity, CPI-455 was used to reduce DTPs in multiple models.426 CPI-455 is more effective in temozolomide (TMZ)-resistant glioblastoma cells than in TMZ-native cells. Thus, CPI-455 may be a sensitizing agent for TMZ in glioblastoma, indicating the combinational potential of targeting the epigenetic landscape with cytotoxic therapies.429 Likewise in leukemia, CPI-455 treatment sensitized acute promyelocytic leukemia (APL) cells to all-trans retinoic acid-induced differentiation.430 Importantly, CPI-455 may also be applied in the context of inflammatory diseases. Dexmedetomidine (DEX) is frequently used to prevent excessive inflammatory response in sepsis-induced organ failure. In a mouse model with acute kidney injury, DEX and CPI-455 achieved the same effect in decreasing H3K4me3 enrichment of multiple inflammatory cytokine genes. Thus, DEX can be used to attenuate acute kidney injury by blocking JARID1 during sepsis.431

Another JARID1 inhibitor KDOAM-25, has a half maximal inhibitory concentration of <100 nM for JARID1A-D, and demonstrates no off-target effect on a panel of 55 other enzymes. As discussed earlier, JARID1B is an oncogenic factor for multiple myeloma. The treatment of multiple myeloma cells with KDOAM-25M led to decreased cell proliferation and increased global H3K4 methylation level at transcription sites.432 Furthermore, ryuvidine might be a lead compound for JARID1-targeting therapeutics, which exerted its inhibitory activity on not only JARID1A but also recombinant JARID1B and C.433 Recent research assessed the antitumor effect of KDM5-inh1, a novel panel of selective JARID1 inhibitors in multiple cancer cell lines, and found that JARID1 inhibition is especially effective in HER2 + breast cancer, which might serve as a diagnostic tool for the selection of target patients.434

Conclusion and future perspectives

JMJD protein family members regulate multiple tumor-associated genes either dependent or independent of its histone demethylase activity according to different cellular contexts. Growing evidence has suggested the diverse functions of JMJD class of histone demethylase in pathological processes, justifying the development of small-molecule inhibitors against JMJD proteins. The utility of JMJD protein inhibitors should not be limited to combating cancer but also the treatment of inflammatory disorders such as asthma. However, several obstacles need to be overcome in the application of JMJD proteins as treatment targets for cancer and inflammatory diseases.

First, as some of them promote cancer progression, their overexpression or the activation of their enzymatic functions could be hallmarks of tumorigenesis. Nevertheless, the cancer-suppressive activities have also been implicated for some JMJD family members such as JMJD3 which exhibits a dual role in CRC progression, making it both a tumor suppressor and a tumor activator. Thus, before using JMJD protein blockade for treatment, it is important to elucidate the specific functions of each JMJD protein in cancer under different conditions.

Secondly, due to its highly polar 2-OG binding pocket, the development of small-molecule inhibitors for the JMJD family has lagged behind, with several JMJD inhibitors being reported but functionally inactive. Moreover, despite growing research, no known inhibitors to date are commercially available for the treatment of any cancer type. For instance, various JMJD2 inhibitors have been reported as cancer therapeutic agents, but currently there is only one agent under clinical evaluation. To eradicate non-selective target effects and improve the selectivity of JMJD inhibitors, further studies on the structural information and structure-activity relationship of JMJD proteins are warranted.