Next Article in Journal
Synthesis, Physicochemical, Labeling and In Vivo Characterization of 44Sc-Labeled DO3AM-NI as a Hypoxia-Sensitive PET Probe
Next Article in Special Issue
Mechanistic Analysis of Chemically Diverse Bromodomain-4 Inhibitors Using Balanced QSAR Analysis and Supported by X-ray Resolved Crystal Structures
Previous Article in Journal
Dihydroconiferyl Ferulate Isolated from Dendropanax morbiferus H.Lév. Suppresses Stemness of Breast Cancer Cells via Nuclear EGFR/c-Myc Signaling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bromodomain and Extra-Terminal Protein Inhibitors: Biologic Insights and Therapeutic Potential in Pediatric Brain Tumors

1
Department of Pediatric Oncology, Dana-Farber Boston Children’s Cancer and Blood Disorders Center, Boston, MA 02215, USA
2
Broad Institute of Harvard and MIT, Cambridge, MA 02142, USA
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2022, 15(6), 665; https://doi.org/10.3390/ph15060665
Submission received: 29 April 2022 / Revised: 19 May 2022 / Accepted: 23 May 2022 / Published: 26 May 2022
(This article belongs to the Special Issue Bromodomains: A New Target Class for Drug Development)

Abstract

:
Pediatric brain tumors have surpassed leukemia as the leading cause of cancer-related death in children. Several landmark studies from the last two decades have shown that many pediatric brain tumors are driven by epigenetic dysregulation within specific developmental contexts. One of the major determinants of epigenetic control is the histone code, which is orchestrated by a number of enzymes categorized as writers, erasers, and readers. Bromodomain and extra-terminal (BET) proteins are reader proteins that bind to acetylated lysines in histone tails and play a crucial role in regulating gene transcription. BET inhibitors have shown efficacy in a wide range of cancers, and a number have progressed to clinical phase testing. Here, we review the evidence for BET inhibitors in pediatric brain tumor experimental models, as well as their translational potential.

Graphical Abstract

1. Introduction

Pediatric brain tumors are the most common solid tumor in children, and are the leading cause of cancer-related death [1]. Several aggressive pediatric brain tumors, such as medulloblastoma (MB), high-grade glioma (HGG), ependymoma (EPN) and atypical teratoid/rhabdoid tumors (ATRTs), are defined by epigenetic dysregulation that has been shown to reflect disordered developmental processes that occur in susceptible cell types and along specific spatio-temporal patterns [2,3]. In the context of normal development, epigenetic mechanisms allow for transcriptional control independent of DNA sequence, and are crucial components of cell differentiation and specialization [4,5,6]. The most widely studied mechanisms include DNA methylation and histone modifications, both of which are commonly co-opted in pediatric brain tumor pathogenesis.
Histones are the fundamental building blocks of chromatin and assemble in octamers, around which DNA is wrapped to form the nucleosome [7]. Each histone has an amino acid tail that extends from the nucleosome and is enriched in positively charged residues, such as lysine and arginine, that are subject to post-translational covalent modifications that have dramatic impacts on chromatin accessibility [8,9]. At least 16 of these modifications have been described to date and include lysine acetylation, lysine and arginine methylation, serine and threonine phosphorylation, and lysine ubiquitination or sumoylation [10,11]. Specific enzymes, known as “writers” and “erasers”, are responsible for catalyzing the transfer of these marks, while “reader” proteins are involved in recognition of the marks to recruit other protein complexes involved in transcriptional control.
Histone lysine acetylation is one of the most well-studied histone post-translational modification (PTM) and decreases the positive charge of histones leading to a relaxed chromatin conformation associated with greater levels of transcription [12,13]. Histone acetylation writers (histone acetyltransferases, or HATs), erasers (histone deacetylases, or HDACs) and readers (bromodomains, BRDs) have been well characterized and each has been implicated in the pathogenesis of a wide range of cancers [14,15,16,17,18,19]. Over 60 bromodomains have been identified in humans, occurring in over 40 proteins (each containing between 1–6 BRDs) [20]. One important class of bromodomains are the bromodomain and extra terminal (BET) family, that is made up of BRD2, BRD3, BRD4, and BRDT [21]. Each BET protein is characterized by two tandem bromodomains (BD1 and BD2) in their N-terminal region and a C-terminal extra-terminal domain [22,23]. In particular, BRD4 plays a key role in RNA polymerase II dependent transcription through the recruitment of the positive transcription elongation factor complex (P-TEFb) and the Mediator complex to promoter regions [24,25]. Through this activity, BRD4 serves as a critical transcriptional coactivator at regions of hyperacetylation (e.g., enhancers or super-enhancers) and mediates expression of key transcription factors, such as c-MYC [26,27]. Since the discovery of small molecule BET inhibitors in 2010, their activity has been demonstrated in a wide range of transcriptionally addicted human cancers. Here we will review the rationale and translational potential for BET inhibitors in pediatric brain cancers.

2. BET Protein Structure and Function

There are four BET proteins, which in humans are referred to as bromodomain-containing protein 2 (BRD2), bromodomain-containing protein 3 (BRD3), bromodomain-containing protein 4 (BRD4), and bromodomain testis associated protein (BRDT). BRD2/3/4 are all ubiquitously expressed, while BRDT is limited to male germ cell tissue. In central nervous system (CNS) tissue, specifically, BRD2/3/4 have all been found to be highly expressed, although expression levels can vary depending on cell-type and spatial location (Figure 1A). BET proteins (particularly BRD2 and BRD4) have also been found to be highly expressed in the developing mouse brain (Figure 1B). Inhibition, via genetic knockout or inhibitor treatment, has been linked with diverse phenotypes, including cerebellar ataxia, seizures, and autism-like behaviors [28,29,30].
BET proteins are comprised of two N-terminal bromodomains (BD1, BD2), an extraterminal domain (ET), and a C-terminal domain (CTD). Each bromodomain contains four alpha helices separated by a variable loop region. This forms a hydrophobic pocket that anchors to acetylated lysine residues via a conserved asparagine residue. The BD1/BD2 amino acid residues critical for acetylated lysine binding are highly conserved across BET proteins; however, there are substantial differences between BD1 and BD2 active sites, which confer functional differences and the ability to chemically target selectively [31]. The BET extra-terminal domain (ET) is able to recruit other chromatin-regulating proteins, such as NSD3, JMJDs, and CHD4 [32]. The C-terminal domain (CTD) is present only in BRD4 and BRDT, and is responsible for recruitment of the positive elongation factor (P-TEFb) [33]. BET proteins utilize these unique structural elements to read acetylated histones at cis-regulatory elements, such as promoters and enhancers, and serve as transcriptional coactivators. Indeed, one of the main mechanisms by which BET inhibitors exhibit anti-cancer effects is through targeting super-enhancer driven oncogenes, such as c-MYC [27].
Figure 1. BET protein expression in the central nervous system. (A) Human bulk RNA expression from the GTex database, clustered by brain tissue type [34] (B) Single-cell RNA sequencing atlas of the developing mouse brain showing ubiquitous BRD4 expression throughout cell types of the developing central nervous system [35,36].
Figure 1. BET protein expression in the central nervous system. (A) Human bulk RNA expression from the GTex database, clustered by brain tissue type [34] (B) Single-cell RNA sequencing atlas of the developing mouse brain showing ubiquitous BRD4 expression throughout cell types of the developing central nervous system [35,36].
Pharmaceuticals 15 00665 g001

3. BET Inhibitors

Small molecule inhibitors of BET bromodomains were first discovered in 2010 by two groups, based on a thienotriazolodiazepine (JQ1) and benzotriazolodiazepine (I-BET151) scaffold, respectively [37,38]. They were shown to potently inhibit bromodomain binding to acetylated lysine residues, resulting in displacement of BRD4 from nuclear chromatin. Both compounds selectively inhibit BET bromodomains over other bromodomain classes, although neither discriminates between BD1/BD2 within each BET protein nor between BRD2/BRD3/BRD4/BRDT. These compounds have demonstrated impressive pre-clinical efficacy in a variety of cancer models, including BRD4–NUT fusions NUT-midline carcinoma (NMC), acute myeloid leukemia (AML), medulloblastoma, breast cancer, and lung cancer [39,40,41]. In the last decade, several “clinical-grade” pan-BET inhibitors with improved pharmacokinetic profiles have been developed (Table 1). From the initial diazepane-based JQ1 and I-BET151, a variety of scaffolds with potent BET inhibitory properties have been employed successfully. Clinical trials with these agents have shown that thrombocytopenia is an important and common dose-limiting toxicity of pan-BET inhibitors [42].
More recently, BD2 selective BET inhibitors, such as ABBV-744, have been developed with the goal to narrow anti-neoplastic selectivity and reduce off-target hematologic toxicities [43,44]. Another exciting development has been in the field of targeted protein degradation. In 2015, three BET proteolysis-targeting chimeras (PROTACs) were discovered: dBET1, MZ1, and ARV-825 [45,46,47]. Proteolysis-targeting chimeras (PROTACs) are bifunctional small molecules that include a “warhead” ligand that recruits a protein of interest, linked to an E3 ligase ligand, that induces target ubiquitination and subsequent proteasomal degradation. In some settings these compounds have shown enhanced potency in cancer cell lines compared to their parental inhibitors, paving a new strategy for targeting BET proteins in select malignancies.

4. BET Inhibitors in Pediatric Brain Tumor Models

4.1. Medulloblastoma

Medulloblastoma is the most common embryonal brain tumor occurring in childhood, and accounts for approximately 6.5% of all pediatric brain tumors. Several landmark studies, published in 2006-2011, used genomic profiling to identify distinct molecular subgroups, which were summarized in a 2012 consensus report designating four subgroups (WNT, SHH, Group 3, Group 4) [48,49,50,51,52]. In 2017, three independent groups also used DNA methylation analysis, which recently has been demonstrated to have exquisite discriminatory power in the diagnosis of CNS tumors, to further subclassify medulloblastoma subtypes [53,54,55]. Each medulloblastoma subgroup has unique underlying biology, which is reflected in differences in patient outcomes and treatment responses. These subgroups have subsequently been incorporated into the 2016 and 2021 updated World Health Organization (WHO) Classification of Tumors of the Central Nervous System [56,57].
Treatment for medulloblastoma is aggressive and includes surgical resection, radiation therapy, and chemotherapy. While these treatments have led to five-year overall survival rates over 80%, they can lead to significant treatment-related toxicities, and patients with higher risk features can have survival rates as low as 20% [58]. One subgroup that is associated with a poor prognosis is Group 3 medulloblastoma, which are often metastatic at diagnosis and tend to be refractory to therapy or have late recurrence. Molecularly, Group 3 tumors often have MYC amplification or high MYC expression, and high MYC expression has been reported to be an independent risk factor for poor outcome in medulloblastoma [59].
Several groups have reported promising findings of BET inhibition in MYC-amplified medulloblastoma. The BET inhibitor JQ1 potently decreases cell viability of MYC-amplified medulloblastoma cell lines, causes G1 arrest and apoptosis, and downregulates MYC-targets, as well as MYC transcription itself [60,61]. Furthermore, JQ1 treatment shows in-vivo efficacy with decreased tumor growth and prolonged survival in both flank and cerebellar orthotopic models of Group 3 medulloblastoma (Table 2). JQ1 has also been shown to induce cellular senescence and suppress transcriptional programs associated with poor prognosis in medulloblastoma patients [62].
BET inhibition has also been shown to be a promising strategy in SHH-driven medulloblastoma. Canonical hedgehog (Hh) signaling occurs via binding of Hh ligands to the PTCH1 transmembrane protein, which leads to the cessation of PTCH1′s role in repressing the G protein coupled receptor Smoothened (SMO) [63]. Activation of SMO leads to positive regulation of the GLI zinc-finger transcription factors, leading to the expression of a Hh transcriptional program that regulates proliferation and cell specification [64]. Aberrant Hh signaling has been associated with several cancers, including basal cell carcinoma, medulloblastoma, and pancreatic cancer [65,66]. BRD4 has been shown to be critical in GLI1 and GLI2 transcription via direct promoter occupation using a mouse 3T3 Gli-luciferase reporter cell line [67]. This interaction can be significantly disrupted with JQ1 treatment, providing a biological rationale for BET inhibition in SHH-driven medulloblastoma. Indeed, JQ1 decreases proliferation and viability of SHH-driven medulloblastoma cell lines in-vitro and in-vivo, even when mutations conferring these cell lines to SMO inhibitor resistance are present. The BET inhibitor I-BET151 has also been studied, and was found to significantly attenuate HH activity in the Light2 reporter cell line [68]. Furthermore, I-BET151 decreases growth and viability of a murine derived (Ptch1+/−) medulloblastoma cell line in-vitro, decreases GLI1 transcription, and decreases tumor growth in-vivo.
Several synergistic drug combinations with BET inhibitors have also been investigated, with particular focus on cell cycle inhibitors. In 2019, Bandopadhayay et al. used a functional genomic approach including CRISP/Cas9-mediated loss of function and ORF/cDNA rescue screens to identify key regulators of BET inhibitor response and resistance [69]. They found that resistant cells express transcriptional programs associated with neuronal differentiation, while still maintaining proliferative potential. They also demonstrated that CDK4/CDK6 inhibition delays the acquisition of BET inhibitor resistance, and the combination of JQ1 with LEE01 (a CDK4/6 inhibitor) improved survival, compared to monotherapy in flank and orthotopic xenograft models of MYC-driven medulloblastoma. The CDK2 inhibitor Milcilib also synergizes with JQ1 in Group 3 medulloblastoma in-vivo and in-vivo models [70]. From a mechanism standpoint, combination treatment with JQ1 + Milcilib targets MYC expression, as well as MYC stabilization, respectively; the latter being a known effect of CDK2 inhibition through suppression of MYC residue S62 phosphorylation. Through these mechanisms, combination treatment was found to be significantly more effective than either JQ1 or Milcilib monotherapy in downregulating MYC target expression in Group 3 medulloblastoma models.

4.2. Diffuse Intrinsic Pontine Glioma

Diffuse intrinsic pontine glioma (DIPG), now referred to as diffuse midline glioma, H3 K27-altered in the 2021 WHO classification, is a lethal brainstem tumor occurring in childhood. Radiation is the only proven therapeutic, despite dozens of chemotherapeutic trials over the last decades, and it only confers a survival advantage of approximately three months [71]. In 2012, a breakthrough was made when several groups reported that over 80% of DIPG tumors harbor a unique lysine-to-methionine mutation (K27M), the H3 histone tail [72,73]. Subsequent work showed that this mutation leads to widespread loss of the repressive H3K27me3 mark through inhibition of the poly-comb repressor 2 complex (PRC2) [74]. Importantly, there are some genomic regions where PRC2 activity is retained and even increased [75].
ChIP sequencing studies have shown that in DIPG, the pathogenic H3K27M mutation colocalizes with acetylated H3K27, RNA polymerase II, and BRD2/BRD4 at sites of active transcription [76]. Treatment with JQ1 impairs the growth and viability of DIPG cell lines, with RNA sequencing demonstrating specific targeted effects at genes with BRD2/BRD4 promoter occupancy. Furthermore, 10 days of JQ1 treatment was found to significantly prolong survival in a DIPG mouse orthotopic (brainstem) patient derived xenograft model. Nagaraja et al. found similar promising findings with JQ1 treatment in a panel of H3K27M mutant DIPG cell lines, while a H3WT glioblastoma cell line (SU-pcGBM2) showed little response [77]. Transcriptomic analysis after JQ1 treatment showed downregulation of genes associated with central nervous system development, such as NTRK3, ASCL1, and MYT1. Lentiviral shRNA knockdown of BRD4 in two DIPG models decreased tumor growth and prolonged survival after mouse orthotopic brainstem injections with lentiviral-modified and control cell lines.
Drug combinations with BET inhibitors have also been investigated in DIPG. Wiese et al. examined JQ1 with the CREB binding protein (CBP) inhibitor ICG-001 [78]. CBP’s main function is as an acetyltransferase which regulates H3K27 acetylation, although it also recruits transcription factors and serves a scaffolding role in multi-unit transcriptional complexes [79]. The authors found that JQ1 + ICG-001 treatment synergistically inhibited DIPG proliferation and invasion potential in DIPG cell lines. RNA sequencing showed that JQ1 downregulated a significant proportion of super-enhancer regulated genes, while surprisingly ICG-001 monotherapy led to an upregulation of these genes. Combination therapy was able to reverse this subset of SE-associated genes which were inadvertently increased with ICG-001 treatment. Another study used H3K27M/PDGFB expressing NSCs as a DIPG model, and showed that JQ1 synergized with the EZH2 inhibitor EPZ6438 (tazemetostat) in-vitro and in-vivo [80]. Taylor et al. showed that targeting the NOTCH pathway with the γ-secretase inhibitor MRK003 also synergized with JQ1 in 2/3 DIPG cell lines tested [81].
Recently, high-throughput chromosome conformation capture (Hi-C) has been used to characterize the 3-dimensional chromatin structure of DIPG cells [82]. This technology enables mapping of tumor-specific regulatory networks, as well as enhancer hijacking events. The BET inhibitor BMS986158 and the BET degrader dBET6 were found to significantly perturb the micro- and macro-chromatin interactions of DIPG cells, and these findings were uniformly more pronounced with dBET6 treatment. This study suggests one possible advantage of targeted protein degradation over catalytic inhibition in this context.

4.3. Ependymoma

Ependymoma is a brain tumor that occurs in both children and adults and arises from ependymal cells lining the ventricular system and spinal canal. Ependymoma is the third most common brain tumor in childhood, and can arise anywhere along the craniospinal axis; although in children they tend to occur more commonly in a supratentorial (ST) or posterior fossa (PF) location [83]. Diagnosis of ependymoma is made histologically with three distinct histologic grades (Grade I, II, and III), although prognostic utility is debated given inter-observer variability and the advent of molecular stratification [84]. Standard of care involves maximum safe surgical resection with post-operative radiotherapy [85]. The role of chemotherapy is controversial, although it is commonly used in patients under 18 months of age to delay radiation and there are ongoing efforts to understand if there are benefits to adjuvant and/or maintenance chemotherapy regimens for select patients [86].
In a landmark 2015 study, Patjler et al. used DNA methylation and transcriptomic profiling to identify at least nine clinically and biologically relevant distinct subgroups, which have had major implications in the official WHO classification [87]. Supratentorial ependymomas are made up of ST-sub-ependymoma and two subgroups, defined by characteristic oncogenic fusions (YAP1 and RELA). Of note, RELA ependymomas have been renamed in the most recent WHO classification to ZFTA-fusion positive with the recognition that C11orf95/ZFTA fusions can occur with non-RELA partner genes, such as MAML2/3, NCOA1/2, MN1, or CTNNA2 [88,89,90]. Posterior fossa ependymomas are comprised of PF-sub-ependymoma and two subgroups, denoted group A (PFA, associated with EZHIP overexpression and poor clinical outcome) and group B (PFB, associated with relatively favorable prognosis). Spinal ependymomas are made up of SP-sub-ependymoma, SP myxopapillary ependymoma, and SP ependymoma. A recent study by Bockmayr et al. demonstrated that SP myxopapillary ependymoma is molecularly comprised of two distinct subgroups, MPE-A and MPE-B; with the former being associated with significantly higher relapse rate and decreased progression-free survival [91].
Work by Mack et al. used H3K27ac ChIP-seq to identify enhancers and super-enhancers in a cohort of 42 primary intracranial ependymomas [92]. They found that genetic knockout of super-enhancer regulated genes impaired growth of ependymoma cell lines in-vivo, which provided a rationale for trialing BET inhibition as a therapeutic strategy. They subsequently showed that JQ1 inhibited the growth of two ependymoma cell lines, one PF-A and the other ST-ZFTA. Another group tested the CNS-penetrant BET inhibitor OTX-015 in three ependymoma cell lines (two PFA, one ST (subgroup not specified)) and similarly found it to be effective with sub-micromolar IC50 values in three-day viability assays [93]. They found that in-vivo treatment prolonged survival in 1/2 of the PFA intracranial patient-derived xenograft models tested.

4.4. Embryonal Tumor with Multilayer Rosettes (ETMR)

ETMR is a rare, aggressive CNS embryonal tumor occurring primarily in children younger than three years of age [94]. It was first recognized as a distinct molecular diagnosis after studies identified a unique recurrent microRNA amplification of C19MC on chr19q13.42 in a subset of primitive neuroectodermal tumors, including embryonal tumors, with abundant neuropil and true rosettes (ETANTR), ependymoblastoma, medulloepithelioma and supratentorial primitive neuroectodermal tumors (sPNET) [95,96,97]. Treatment of ETMR involves maximum safe surgical resection, and given the young age at diagnosis craniospinal irradiation is typically contraindicated (although local radiation is often used) [98]. High-dose chemotherapy protocols using agents with known embryonal activity, such as cyclophosphamide, vincristine, methotrexate, etoposide, cisplatin, carboplatin, and thiotepa, are often used [99]. However, given the rarity of the disease no standardized treatment protocol has been established yet. Despite these intensive treatment regimens prognosis remains dismal, with reported five-year survival rates between 0–30%.
In 2019, Sin-Chan et al. comprehensively profiled over 80 primary ETMR samples using global methylation, SNP, transcriptional, and miRNA profiling [100]. Through this analysis they uncovered a C19MC-LIN28A-MYCN feed forward oncogenic circuit. LIN28A had previously been found to be highly expressed in ETMR, and is a pluripotency factor and RNA-binding protein important to neural development. LIN28A has also been implicated in the pathogenesis of many advanced human malignancies, including ovarian carcinoma, germ cell tumors, and Wilms’ tumor [101]. The authors hypothesized that this C19MC-LIN28A-MYCN core regulatory circuit could be targeted by BET protein inhibition, and, indeed, found that treatment with JQ1S (the active isomer of JQ1) led to a reduction in viability of ETMR cell lines, and qRT-PCR and western blot showed downregulation of MYCN and LIN28A.

4.5. Atypical Teratoid/Rhabdoid Tumor (ATRT)

ATRT is a rare, highly aggressive brain tumor of early childhood [102]. Biallelic inactivation of SMARCB1, a core member of the SWI/SNF (also known as BAF) chromatin remodeling complex, is the core genetic driver event in the vast majority of cases [103,104]. While ATRT has been shown to have a “quiet genome” with few other identifiable pathogenic alterations, this genetic simplicity belies the clinical aggressiveness of these tumors and the difficulties in finding effective targeted therapeutics [105]. Recent studies have used transcriptomics and DNA-methylation to group ATRT into three distinct subgroups: ATRT-TYR, ATRT-SHH, and ATRT-MYC [106,107,108].
ATRT expresses high levels of c-MYC, and ChIP-sequencing shows that SMARCB1 loss leads to significant enrichment of MYC chromatin occupancy at transcriptional start sites (TSS), compared to normal embryonic stem cells [109]. Genetic knockdown of c-MYC decreases ATRT cell growth in-vivo and prolongs survival of mouse PDX models in-vivo, providing a rationale for BET inhibition in this malignancy. Indeed, JQ1 treatment significantly decreases c-MYC transcription, as well as MYC-driven stemness programs regulated by SOX2, Nanog and OCT4 in ATRT. Furthermore, JQ1 treatment significantly prolongs survival in orthotopic xenograft mouse models.
Table 2. Findings from pre-clinical studies of BET inhibitors in pediatric brain tumor models.
Table 2. Findings from pre-clinical studies of BET inhibitors in pediatric brain tumor models.
TumorIn-Vivo Cell Line/ModelBET InhibitorNotable FindingsCitation
MBMB002, Group 3 MBJQ1JQ1 was effective in broad panel of MB cell lines, induced apoptosis and G1 cell cycle arrest. RNA sequencing showed decreased MYC and MC-target expression. Orthotopic xenograft (cerebellar) showed increased survival with JQ1 treatmentBandopadhayay, [60]
MBHD-MB3, MYC amplified Group 3 MBJQ1JQ1 effective in broad panel of cell lines, inducing apoptosis and G1 cell cycle arrest. Caused decreased MYC and MYC-targets’ expression, and affected components of p53 and cell cycle pathway. Flank xenograft study showed decreased tumor growth and prolonged survivalHennsen et al. [61]
MBDAOY, MYC-driven MBJQ1JQ1 effective in MB cell lines, induced apoptosis and cell cycle arrest. They also showed that it induced cellular senescence, and that transcriptional programs suppressed by treatment are associated with adverse risk in MB patients. Flank xenograft study showed decreased tumor growthVenkataraman et al. [62]
MBMED1-MB, SMO-WT/SMO-D477G-MB (autochthonous derived from Ptch+/−; Tpr53−/− and Ptch+/−; lacZ mice, respectivelyJQ1JQ1 decreased proliferation and viability of SHH-driven MB in-vitro and in-vivo (flank and cerebellar models used), even when cell lines had SMO inhibitor resistance mutationsTang et al. [67]
MBMurine Ptch+/− MB modelI-BET151I-BET151 decreased SHH-driven MB growth in-vivo, and decreased Gli1 expression. I-BET151 was effective in decreasing tumor growth in-vivo (subcutaneous)Long et al. [68]
MBD458 and MB002, MYC driven MBJQ1 + LEE01CDK4/CDK6 inhibition delayed development of BET inhibitor resistance. Combination of JQ1 with LEE01 (a CDK4/6 inhibitor) improved survival in flank and orthotopic xenograft models of MYC-driven medulloblastomaBandopadhayay, [69]
MBGTML2 (murine derived Group 3 MB) and MB002JQ1 + MilcilibJQ1+ CDK2 inhibitor synergized to induce apoptosis and cell cycle arrest. Combination treatment in-vivo extended survival in two orthoptic (cerebellar) models of Group 3 MBBolin et al. [70]
DIPGSF8628JQ1JQ1 impaired DIPG growth and viability in-vivo, and improved survival in an orthotopic mouse PDX model (brainstem)Piunti et al. [76]
DIPGSU-DIPG-VI and SF7761BRD4 shRNAJQ1 decreased growth of DIPG cell lines and downregulated genes associated with CNS development. Lentiviral shRNA knockdown of BRD4 extended survival of mice bearing two different orthotopic (brainstem) DIPG modelsNagaraja et al. [77]
DIPGN/A (no in-vivo data)JQ1 + ICG-001BET + CBP inhibition synergized to decrease growth and viability of DIPG cell lines, and preferentially downregulated super-enhancer genesWiese et al. [78]
DIPGH3K27M/PDGFB expressing NSCsJQ1 + TazemtostatBET + EZH2 inhibition was a synergistic combination in H3K27M/PDGFB transformed NSCsZhang et al. [80]
DIPGN/A (no in-vivo data)JQ1 + MRK003BET inhibition + NOTCH inhibition synergized in 2/3 DIPG models to induce apoptosis and cell deathTaylor et al. [81]
DIPGN/A (no in-vivo data)BMS986158, dBET6BET inhibition and degradation significantly altered the chromatin architecture of DIPG via Hi-C analysis, although the effect was more pronounced with BET degradationWang et al. [82]
EpendymomaN/A (no in-vivo data)JQ1JQ1 inhibited proliferation and viability of one supratentorial (H.EP1) and one PF-A (H.612) ependymoma cell lineMack et al. [92]
EpendymomaEPP-MI and EPV-FL-MI (PFA)OTX015OTX015 induced apoptosis and cell cycle arrest in two PFA and one ST (subtype not specified) models of ependymoma. In-vivo OTX015 extended survival of the EPP-MI orthotopic intracranial PDX model, but had no improvement in the EPP-FL-MI modelServidei et al. [93]
ETMRN/A (no in-vivo data)JQ1S (active isomer of JQ1)JQ1S decreased growth and viability of ETMR cell lines in-vivo, and downregulated MYCN and LIN28A expressionSin-Chan et al. [100]
ATRTMAF-737JQ1JQ1 potently inhibited viability of ATRT (MYC subtype) cell lines, and decreased transcription of c-MYC targets and c-MYC itself. JQ1 prolonged survival in an orthotopic (cerebellar) ATRT modelAllimova et al. [109]
ATRT = atypical teratoid rhabdoid tumor, DIPG = diffuse intrinsic pontine glioma, ETMR = embryonal tumor with multilayer rosettes, MB = medulloblastoma.

5. BET Inhibitors in the Clinic

5.1. BET Inhibitors in CNS Malignancies

Despite the development of several clinical grade BET inhibitors, few have demonstratable CNS penetrance. While JQ1 is brain penetrant and has widely been used in the pre-clinical setting, it’s poor pharmacokinetic properties (namely short half-life) have precluded translation. One CNS penetrant BET inhibitor that has been well characterized is OTX015/MK-8628/Birabresib, which was shown to be effective in glioblastoma (GBM) pre-clinical models [110]. This led to a Phase IIa trial in patients with recurrent glioblastoma (NCT02296476). Twelve patients were enrolled, and the drug was well tolerated with pharmacokinetic studies demonstrating biologically active levels; however, all patients progressed with a median progression-free survival of two months and the trial was terminated [111]. BMS-986378/CC-90010 is an orally bioavailable, CNS-penetrant BET inhibitor that is currently being evaluated in clinical trials. A phase I trial of CC-90010 in patients with advanced solid tumors and relapsed/refractory non-Hodgkin’s lymphoma enrolled a total of 69 patients, of whom 10 had high-grade gliomas (NCT03220347). There was an overall response rate of 2.9% (n = 2), with 8.8% (n = 6) of patients achieving stable disease [112]. One patient with a progressive grade II diffuse astrocytoma had a complete response (CR). There are two ongoing studies of CC-90010 in high-grade glioma, including a phase I trial to evaluate the CNS penetration of CC-90010 in patients with progressive/recurrent astrocytoma, anaplastic astrocytoma or GBM (NCT04047303), as well as a Phase Ib study of CC-90010 in combination with temozolomide and radiation in patients with newly diagnosed GBM (NCT04324840).

5.2. BET Inhibitors in Pediatrics

In 2019, the first pediatric trial of BET inhibition opened (NCT03936465) investigating the inhibitors BMS-986158 (Arm 1) and BMS-986378/CC-90010 (Arm 2). The primary aims of the trial are to determine toxicities and recommended phase II doses for these agents, while secondary outcomes include: (a) efficacy, (b) pharmacokinetics, and (c) a host of pharmacodynamic and predictive biomarkers. Patients must be <21 and able to swallow intact pills. Each arm has two cohorts, one being with unselected biology (relapsed/refractory solid tumors or lymphoma for Arm 1, CNS tumors for arm 2) and the other enriched for predictive biology. Eligibility for this second criterion includes: MYCN amplification or high copy number gain, MYC amplification or high copy number gain, translocation involving MYC or MYCN, translocation involving BRD4 or BRD3, BRD4 amplification or high copy number gain, and/or histologic diagnosis of NUT midline carcinoma (NMC).
The state of BET inhibitor development in pediatric oncology was recently summarized at a strategy forum organized by ACCELERATE, in collaboration with the European Medicines Agency (EMA) and the US Food and Drug Administration (FDA) [113]. Given that most adult trials of pan-BET inhibitors have been challenged by similar issues (narrow therapeutic index, due to thrombocytopenia, modest anti-tumor activity as monotherapy), it was agreed that opening additional trials to study pan-BET inhibitors, other than BMS-986158 and BMS-986378/CC-90010, was not a worthwhile strategy. Agents with unique properties, such as BDII-selectivity, dual targets e.g., BET/p300, or improved CNS penetrance were highlighted as areas of particular interest worthy of further investigation.

6. Conclusions

The discovery and development of BET inhibitors was a landmark finding in cancer epigenetics, and many drugs in this class have advanced to clinical trial stages. Results to date have mostly been mixed, with modest anti-tumor activity and relatively narrow therapeutic index, due to hematologic dose-limiting toxicities. However, there is strong biological and pre-clinical rationale for testing BET inhibitors in pediatric brain tumors, given the common thread of epigenetic dysregulation. The ongoing trial of BMS-986378/CC-90010 in relapsed/refractory pediatric CNS tumors will be highly informative to assess the promise of BET inhibitors in this patient population. Based on experience from experimental models and adult clinical trials, it is likely that monotherapy alone will not be sufficient and, therefore, continued research dedicated to investigating synergistic combinations with BET inhibitors will be critical.

Author Contributions

Conceptualization, writing, and editing: A.G., J.C. and M.G.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study (Figure 1) is openly available at the Gtex portal [34] and the Mouse Brain Atlas [35,36].

Acknowledgments

We sincerely thank the members of the Filbin and Qi laboratories for their valuable comments. We apologize to authors of studies which, for reasons of text space, could not be referenced within this work. Figures were created with Biorender.com (accessed on 5 February 2022) and Adobe Illustrator.

Conflicts of Interest

All authors declare no financial conflict of interest.

References

  1. Ostrom, Q.T.; Cioffi, G.; Waite, K.; Kruchko, C.; Barnholtz-Sloan, J.S. CBTRUS Statistical Report: Primary Brain and Other Central Nervous System Tumors Diagnosed in the United States in 2014–2018. Neuro Oncol. 2021, 23, iii1–iii105. [Google Scholar] [CrossRef] [PubMed]
  2. Guerreiro Stucklin, A.S.; Ramaswamy, V.; Daniels, C.; Taylor, M.D. Review of molecular classification and treatment implications of pediatric brain tumors. Curr. Opin. Pediatr. 2018, 30, 3–9. [Google Scholar] [CrossRef] [PubMed]
  3. Cruzeiro, G.A.V.; Rota, C.; Hack, O.A.; Segal, R.; Filbin, M.G. Understanding the epigenetic landscape and cellular architecture of childhood brain tumors. Neurochem. Int. 2020, 144, 104940. [Google Scholar] [CrossRef] [PubMed]
  4. Atlasi, Y.; Stunnenberg, H.G. The interplay of epigenetic marks during stem cell differentiation and development. Nat. Rev. Genet. 2017, 18, 643–658. [Google Scholar] [CrossRef] [PubMed]
  5. Allis, C.D.; Jenuwein, T. The molecular hallmarks of epigenetic control. Nat. Rev. Genet. 2016, 17, 487–500. [Google Scholar] [CrossRef]
  6. Surani, M.A.; Hayashi, K.; Hajkova, P. Genetic and Epigenetic Regulators of Pluripotency. Cell 2007, 128, 747–762. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Luger, K.; Mäder, A.W.; Richmond, R.K.; Sargent, D.F.; Richmond, T.J. Crystal structure of the nucleosome core particle at 2.8 Å resolution. Nature 1997, 389, 251–260. [Google Scholar] [CrossRef]
  8. Allfrey, V.G.; Faulkner, R.; Mirsky, A.E. Acetylation and Methylation of Histones and Their Possible Role in the Regulation of Rna Synthesis. Proc. Natl. Acad. Sci. USA 1964, 51, 786–794. [Google Scholar] [CrossRef] [Green Version]
  9. Khorasanizadeh, S. The Nucleosome: From Genomic Organization to Genomic Regulation. Cell 2004, 116, 259–272. [Google Scholar] [CrossRef] [Green Version]
  10. Zhao, Z.; Shilatifard, A. Epigenetic modifications of histones in cancer. Genome Biol. 2019, 20, 245. [Google Scholar] [CrossRef]
  11. Kouzarides, T. Chromatin modifications and their function. Cell 2007, 128, 693–705. [Google Scholar] [CrossRef] [Green Version]
  12. Hebbes, T.R.; Thorne, A.W.; Crane-Robinson, C. A direct link between core histone acetylation and transcriptionally active chromatin. EMBO J. 1988, 7, 1395–1402. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, Z.; Zang, C.; Rosenfeld, J.A.; Schones, D.E.; Barski, A.; Cuddapah, S.; Cui, K.; Roh, T.-Y.; Peng, W.; Zhang, M.Q.; et al. Combinatorial patterns of histone acetylations and methylations in the human genome. Nat. Genet. 2008, 40, 897–903. [Google Scholar] [CrossRef] [Green Version]
  14. Brownell, J.E.; Zhou, J.; Ranalli, T.; Kobayashi, R.; Edmondson, D.G.; Roth, S.Y.; Allis, C.D. Tetrahymena Histone Acetyltransferase A: A Homolog to Yeast Gcn5p Linking Histone Acetylation to Gene Activation. Cell 1996, 84, 843–851. [Google Scholar] [CrossRef] [Green Version]
  15. Kleff, S.; Andrulis, E.D.; Anderson, C.W.; Sternglanz, R. Identification of a Gene Encoding a Yeast Histone H4 Acetyltransferase. J. Biol. Chem. 1995, 270, 24674–24677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Parthun, M.R.; Widom, J.; Gottschling, D.E. The Major Cytoplasmic Histone Acetyltransferase in Yeast: Links to Chromatin Replication and Histone Metabolism. Cell 1996, 87, 85–94. [Google Scholar] [CrossRef] [Green Version]
  17. Taunton, J.; Hassig, C.A.; Schreiber, S.L. A Mammalian Histone Deacetylase Related to the Yeast Transcriptional Regulator Rpd3p. Science 1996, 272, 408–411. [Google Scholar] [CrossRef]
  18. Dhalluin, C.; Carlson, J.E.; Zeng, L.; He, C.; Aggarwal, A.K.; Zhou, M.-M. Structure and ligand of a histone acetyltransferase bromodomain. Nature 1999, 399, 491–496. [Google Scholar] [CrossRef]
  19. Audia, J.E.; Campbell, R.M. Histone Modifications and Cancer. Cold Spring Harb. Perspect. Biol. 2016, 8, a019521. [Google Scholar] [CrossRef]
  20. Filippakopoulos, P.; Picaud, S.; Mangos, M.; Keates, T.; Lambert, J.-P.; Barsyte-Lovejoy, D.; Felletar, I.; Volkmer, R.; Müller, S.; Pawson, T.; et al. Histone Recognition and Large-Scale Structural Analysis of the Human Bromodomain Family. Cell 2012, 149, 214–231. [Google Scholar] [CrossRef] [Green Version]
  21. Taniguchi, Y. The Bromodomain and Extra-Terminal Domain (BET) Family: Functional Anatomy of BET Paralogous Proteins. Int. J. Mol. Sci. 2016, 17, 1849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Huang, B.; Yang, X.-D.; Zhou, M.-M.; Ozato, K.; Chen, L.-F. Brd4 Coactivates Transcriptional Activation of NF-κB via Specific Binding to Acetylated RelA. Mol. Cell. Biol. 2009, 29, 1375–1387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Shi, J.; Vakoc, C.R. The Mechanisms behind the Therapeutic Activity of BET Bromodomain Inhibition. Mol. Cell 2014, 54, 728–736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Liu, W.; Ma, Q.; Wong, K.; Li, W.; Ohgi, K.; Zhang, J.; Aggarwal, A.K.; Rosenfeld, M.G. Brd4 and JMJD6-Associated Anti-Pause Enhancers in Regulation of Transcriptional Pause Release. Cell 2013, 155, 1581–1595. [Google Scholar] [CrossRef] [Green Version]
  25. Jang, M.K.; Mochizuki, K.; Zhou, M.; Jeong, H.-S.; Brady, J.N.; Ozato, K. The Bromodomain Protein Brd4 Is a Positive Regulatory Component of P-TEFb and Stimulates RNA Polymerase II-Dependent Transcription. Mol. Cell 2005, 19, 523–534. [Google Scholar] [CrossRef]
  26. Delmore, J.E.; Issa, G.C.; Lemieux, M.E.; Rahl, P.B.; Shi, J.; Jacobs, H.M.; Kastritis, E.; Gilpatrick, T.; Paranal, R.M.; Qi, J.; et al. BET Bromodomain Inhibition as a Therapeutic Strategy to Target c-Myc. Cell 2011, 146, 904–917. [Google Scholar] [CrossRef] [Green Version]
  27. Lovén, J.; Hoke, H.A.; Lin, C.Y.; Lau, A.; Orlando, D.A.; Vakoc, C.R.; Bradner, J.E.; Lee, T.I.; Young, R.A. Selective Inhibition of Tumor Oncogenes by Disruption of Super-Enhancers. Cell 2013, 153, 320–334. [Google Scholar] [CrossRef] [Green Version]
  28. Penas, C.; Maloof, M.E.; Stathias, V.; Long, J.; Tan, S.K.; Mier, J.; Fang, Y.; Valdes, C.; Rodriguez-Blanco, J.; Chiang, C.-M.; et al. Time series modeling of cell cycle exit identifies Brd4 dependent regulation of cerebellar neurogenesis. Nat. Commun. 2019, 10, 3028. [Google Scholar] [CrossRef]
  29. Velíšek, L.; Shang, E.; Velíšková, J.; Chachua, T.; Macchiarulo, S.; Maglakelidze, G.; Wolgemuth, D.J.; Greenberg, D.A. GABAergic Neuron Deficit as an Idiopathic Generalized Epilepsy Mechanism: The Role of BRD2 Haploinsufficiency in Juvenile Myoclonic Epilepsy. PLoS ONE 2011, 6, e23656. [Google Scholar] [CrossRef]
  30. Sullivan, J.M.; Badimon, A.; Schaefer, U.; Ayata, P.; Gray, J.; Chung, C.-W.; Von Schimmelmann, M.; Zhang, F.; Garton, N.; Smithers, N.; et al. Autism-like syndrome is induced by pharmacological suppression of BET proteins in young mice. J. Exp. Med. 2015, 212, 1771–1781. [Google Scholar] [CrossRef] [Green Version]
  31. Cheng, C.; Diao, H.; Zhang, F.; Wang, Y.; Wang, K.; Wu, R. Deciphering the mechanisms of selective inhibition for the tandem BD1/BD2 in the BET-bromodomain family. Phys. Chem. Chem. Phys. 2017, 19, 23934–23941. [Google Scholar] [CrossRef] [PubMed]
  32. Rahman, S.; Sowa, M.E.; Ottinger, M.; Smith, J.A.; Shi, Y.; Harper, J.W.; Howley, P.M. The Brd4 Extraterminal Domain Confers Transcription Activation Independent of pTEFb by Recruiting Multiple Proteins, Including NSD3. Mol. Cell. Biol. 2011, 31, 2641–2652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Yang, Z.; Yik, J.H.; Chen, R.; He, N.; Jang, M.K.; Ozato, K.; Zhou, Q. Recruitment of P-TEFb for stimulation of transcriptional elongation by the bromodomain protein Brd4. Mol Cell 2005, 19, 535–545. [Google Scholar] [CrossRef] [PubMed]
  34. GTex Portal. Available online: http://gtexportal.org/home/ (accessed on 9 March 2022).
  35. La Manno, G.; Siletti, K.; Furlan, A.; Gyllborg, D.; Vinsland, E.; Albiach, A.M.; Langseth, C.M.; Khven, I.; Lederer, A.R.; Dratva, L.M.; et al. Molecular architecture of the developing mouse brain. Nature 2021, 596, 92–96. [Google Scholar] [CrossRef]
  36. Mouse Brain Atlas. Available online: http://mousebrain.org/ (accessed on 12 March 2022).
  37. Filippakopoulos, P.; Qi, J.; Picaud, S.; Shen, Y.; Smith, W.B.; Fedorov, O.; Morse, E.M.; Keates, T.; Hickman, T.T.; Felletar, I.; et al. Selective inhibition of BET bromodomains. Nature 2010, 468, 1067–1073. [Google Scholar] [CrossRef] [Green Version]
  38. Nicodeme, E.; Jeffrey, K.L.; Schaefer, U.; Beinke, S.; Dewell, S.; Chung, C.-W.; Chandwani, R.; Marazzi, I.; Wilson, P.; Coste, H.; et al. Suppression of inflammation by a synthetic histone mimic. Nature 2010, 468, 1119–1123. [Google Scholar] [CrossRef]
  39. Zuber, J.; Shi, J.; Wang, E.; Rappaport, A.R.; Herrmann, H.; Sison, E.A.; Magoon, D.; Qi, J.; Blatt, K.; Wunderlich, M.; et al. RNAi screen identifies Brd4 as a therapeutic target in acute myeloid leukaemia. Nature 2011, 478, 524–528. [Google Scholar] [CrossRef] [Green Version]
  40. Shu, S.; Lin, C.Y.; He, H.H.; Witwicki, R.M.; Tabassum, D.P.; Roberts, J.M.; Janiszewska, M.; Huh, S.J.; Liang, Y.; Ryan, J.; et al. Response and resistance to BET bromodomain inhibitors in triple-negative breast cancer. Nature 2016, 529, 413–417. [Google Scholar] [CrossRef] [Green Version]
  41. Shimamura, T.; Chen, Z.; Soucheray, M.; Carretero, J.; Kikuchi, E.; Tchaicha, J.H.; Gao, Y.; Cheng, K.A.; Cohoon, T.J.; Qi, J.; et al. Efficacy of BET Bromodomain Inhibition in Kras-Mutant Non–Small Cell Lung Cancer. Clin. Cancer Res. 2013, 19, 6183–6192. [Google Scholar] [CrossRef] [Green Version]
  42. Sun, Y.; Han, J.; Wang, Z.; Li, X.; Sun, Y.; Hu, Z. Safety and Efficacy of Bromodomain and Extra-Terminal Inhibitors for the Treatment of Hematological Malignancies and Solid Tumors: A Systematic Study of Clinical Trials. Front. Pharmacol. 2020, 11, 621093. [Google Scholar] [CrossRef]
  43. Faivre, E.J.; McDaniel, K.F.; Albert, D.H.; Mantena, S.R.; Plotnik, J.P.; Wilcox, D.; Zhang, L.; Bui, M.H.; Sheppard, G.S.; Wang, L.; et al. Selective inhibition of the BD2 bromodomain of BET proteins in prostate cancer. Nature 2020, 578, 306–310. [Google Scholar] [CrossRef] [PubMed]
  44. Sheppard, G.S.; Wang, L.; Fidanze, S.D.; Hasvold, L.A.; Liu, D.; Pratt, J.K.; Park, C.H.; Longenecker, K.L.; Qiu, W.; Torrent, M.; et al. Discovery of N-Ethyl-4-[2-(4-fluoro-2,6-dimethyl-phenoxy)-5-(1-hydroxy-1-methyl-ethyl)phenyl]-6-methyl-7-oxo-1H-pyrrolo[2,3-c]pyridine-2-carboxamide (ABBV-744), a BET Bromodomain Inhibitor with Selectivity for the Second Bromodomain. J. Med. Chem. 2020, 63, 5585–5623. [Google Scholar] [CrossRef]
  45. Winter, G.E.; Buckley, D.L.; Paulk, J.; Roberts, J.M.; Souza, A.; Dhe-Paganon, S.; Bradner, J.E. Phthalimide conjugation as a strategy for in vivo target protein degradation. Science 2015, 348, 1376–1381. [Google Scholar] [CrossRef] [Green Version]
  46. Lu, J.; Qian, Y.; Altieri, M.; Dong, H.; Wang, J.; Raina, K.; Hines, J.; Winkler, J.D.; Crew, A.P.; Coleman, K.; et al. Hijacking the E3 Ubiquitin Ligase Cereblon to Efficiently Target BRD4. Chem. Biol. 2015, 22, 755–763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Zengerle, M.; Chan, K.H.; Ciulli, A. Selective Small Molecule Induced Degradation of the BET Bromodomain Protein BRD4. ACS Chem. Biol. 2015, 10, 1770–1777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Cho, Y.-J.; Tsherniak, A.; Tamayo, P.; Santagata, S.; Ligon, A.; Greulich, H.; Berhoukim, R.; Amani, V.; Goumnerova, L.; Eberhart, C.G.; et al. Integrative Genomic Analysis of Medulloblastoma Identifies a Molecular Subgroup That Drives Poor Clinical Outcome. J. Clin. Oncol. 2011, 29, 1424–1430. [Google Scholar] [CrossRef] [PubMed]
  49. Kool, M.; Koster, J.; Bunt, J.; Hasselt, N.E.; Lakeman, A.; Van Sluis, P.; Troost, D.; Meeteren, N.S.-V.; Caron, H.N.; Cloos, J.; et al. Integrated Genomics Identifies Five Medulloblastoma Subtypes with Distinct Genetic Profiles, Pathway Signatures and Clinicopathological Features. PLoS ONE 2008, 3, e3088. [Google Scholar] [CrossRef]
  50. Thompson, M.C.; Fuller, C.; Hogg, T.L.; Dalton, J.; Finkelstein, D.; Lau, C.C.; Chintagumpala, M.; Adesina, A.; Ashley, D.M.; Kellie, S.J.; et al. Genomics Identifies Medulloblastoma Subgroups That Are Enriched for Specific Genetic Alterations. J. Clin. Oncol. 2006, 24, 1924–1931. [Google Scholar] [CrossRef]
  51. Northcott, P.A.; Korshunov, A.; Witt, H.; Hielscher, T.; Eberhart, C.G.; Mack, S.C.; Bouffet, E.; Clifford, S.C.; Hawkins, C.E.; French, P.; et al. Medulloblastoma Comprises Four Distinct Molecular Variants. J. Clin. Oncol. 2011, 29, 1408–1414. [Google Scholar] [CrossRef]
  52. Taylor, M.D.; Northcott, P.A.; Korshunov, A.; Remke, M.; Cho, Y.-J.; Clifford, S.C.; Eberhart, C.G.; Parsons, D.W.; Rutkowski, S.; Gajjar, A.; et al. Molecular subgroups of medulloblastoma: The current consensus. Acta Neuropathol. 2011, 123, 465–472. [Google Scholar] [CrossRef] [Green Version]
  53. Cavalli, F.M.; Remke, M.; Rampasek, L.; Peacock, J.; Shih, D.J.H.; Luu, B.; Garzia, L.; Torchia, J.; Nor, C.; Morrissy, S.; et al. Intertumoral Heterogeneity within Medulloblastoma Subgroups. Cancer Cell 2017, 31, 737–754.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Northcott, P.A.; Buchhalter, I.; Morrissy, A.S.; Hovestadt, V.; Weischenfeldt, J.; Ehrenberger, T.; Gröbner, S.; Segura-Wang, M.; Zichner, T.; Rudneva, V.A.; et al. The whole-genome landscape of medulloblastoma subtypes. Nature 2017, 547, 311–317. [Google Scholar] [CrossRef] [Green Version]
  55. Schwalbe, E.; Lindsey, J.C.; Nakjang, S.; Crosier, S.; Smith, A.J.; Hicks, D.; Rafiee, G.; Hill, R.M.; Iliasova, A.; Stone, T.; et al. Novel molecular subgroups for clinical classification and outcome prediction in childhood medulloblastoma: A cohort study. Lancet Oncol. 2017, 18, 958–971. [Google Scholar] [CrossRef] [Green Version]
  56. Louis, D.N.; Perry, A.; Reifenberger, G.; Von Deimling, A.; Figarella-Branger, D.; Cavenee, W.K.; Ohgaki, H.; Wiestler, O.D.; Kleihues, P.; Ellison, D.W. The 2016 World Health Organization Classification of Tumors of the Central Nervous System: A summary. Acta Neuropathol. 2016, 131, 803–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Louis, D.N.; Perry, A.; Wesseling, P.; Brat, D.J.; Cree, I.A.; Figarella-Branger, D.; Hawkins, C.; Ng, H.K.; Pfister, S.M.; Reifenberger, G.; et al. The 2021 WHO Classification of Tumors of the Central Nervous System: A summary. Neuro Oncol. 2021, 23, 1231–1251. [Google Scholar] [CrossRef]
  58. Juraschka, K.; Taylor, M.D. Medulloblastoma in the age of molecular subgroups: A review. J. Neurosurg. Pediatr. 2019, 24, 353–363. [Google Scholar] [CrossRef] [Green Version]
  59. Ellison, D.W.; Kocak, M.; Dalton, J.; Megahed, H.; Lusher, M.E.; Ryan, S.L.; Zhao, W.; Nicholson, S.L.; Taylor, R.E.; Bailey, S.; et al. Definition of Disease-Risk Stratification Groups in Childhood Medulloblastoma Using Combined Clinical, Pathologic, and Molecular Variables. J. Clin. Oncol. 2011, 29, 1400–1407. [Google Scholar] [CrossRef] [Green Version]
  60. Bandopadhayay, P.; Bergthold, G.; Nguyen, B.; Schubert, S.; Gholamin, S.; Tang, Y.; Bolin, S.; Schumacher, S.E.; Zeid, R.; Masoud, S.; et al. BET Bromodomain Inhibition of MYC-Amplified Medulloblastoma. Clin. Cancer Res. 2013, 20, 912–925. [Google Scholar] [CrossRef] [Green Version]
  61. Henssen, A.; Thor, T.; Odersky, A.; Heukamp, L.; El-Hindy, N.; Beckers, A.; Slpeleman, F.; Althoff, K.; Schäfers, S.; Schramm, A.; et al. BET bromodomain protein inhibition is a therapeutic option for medulloblastoma. Oncotarget 2013, 4, 2080–2095. [Google Scholar] [CrossRef] [Green Version]
  62. Venkataraman, S.; Alimova, I.; Balakrishnan, I.; Harris, P.; Birks, D.K.; Griesinger, A.; Amani, V.; Cristiano, B.; Remke, M.; Taylor, M.; et al. Inhibition of BRD4 attenuates tumor cell self-renewal and suppresses stem cell signaling in MYC driven medulloblastoma. Oncotarget 2014, 5, 2355–2371. [Google Scholar] [CrossRef] [Green Version]
  63. Alcedo, J.; Ayzenzon, M.; Von Ohlen, T.; Noll, M.; Hooper, J.E. The Drosophila smoothened Gene Encodes a Seven-Pass Membrane Protein, a Putative Receptor for the Hedgehog Signal. Cell 1996, 86, 221–232. [Google Scholar] [CrossRef] [Green Version]
  64. Oliver, T.; Grasfeder, L.L.; Carroll, A.L.; Kaiser, C.; Gillingham, C.L.; Lin, S.M.; Wickramasinghe, R.; Scott, M.P.; Wechsler-Reya, R.J. Transcriptional profiling of the Sonic hedgehog response: A critical role for N-myc in proliferation of neuronal precursors. Proc. Natl. Acad. Sci. USA 2003, 100, 7331–7336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Johnson, R.L.; Rothman, A.L.; Xie, J.; Goodrich, L.V.; Bare, J.W.; Bonifas, J.M.; Quinn, A.G.; Myers, R.M.; Cox, D.R.; Epstein, E.H., Jr.; et al. Human Homolog of patched, a Candidate Gene for the Basal Cell Nevus Syndrome. Science 1996, 272, 1668–1671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Thayer, S.P.; di Magliano, M.P.; Heiser, P.W.; Nielsen, C.M.; Roberts, D.J.; Lauwers, G.Y.; Qi, Y.P.; Gysin, S.; Castillo, C.F.-D.; Yajnik, V.; et al. Hedgehog is an early and late mediator of pancreatic cancer tumorigenesis. Nature 2003, 425, 851–856. [Google Scholar] [CrossRef] [Green Version]
  67. Tang, Y.; Gholamin, S.; Schubert, S.; Willardson, M.I.; Lee, A.; Bandopadhayay, P.; Bergthold, G.; Masoud, S.; Nguyen, B.; Vue, N.; et al. Epigenetic targeting of Hedgehog pathway transcriptional output through BET bromodomain inhibition. Nat. Med. 2014, 20, 732–740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Long, J.; Li, B.; Blanco, J.R.; Pastori, C.; Volmar, C.-H.; Wahlestedt, C.; Capobianco, A.; Bai, F.; Pei, X.-H.; Ayad, N.G.; et al. The BET Bromodomain Inhibitor I-BET151 Acts Downstream of Smoothened Protein to Abrogate the Growth of Hedgehog Protein-driven Cancers. J. Biol. Chem. 2014, 289, 35494–35502. [Google Scholar] [CrossRef] [Green Version]
  69. Bandopadhayay, P.; Piccioni, F.; O’Rourke, R.; Ho, P.; Gonzalez, E.M.; Buchan, G.; Qian, K.; Gionet, G.; Girard, E.; Coxon, M.; et al. Neuronal differentiation and cell-cycle programs mediate response to BET-bromodomain inhibition in MYC-driven medulloblastoma. Nat. Commun. 2019, 10, 2400. [Google Scholar] [CrossRef]
  70. Bolin, S.; Borgenvik, A.; Persson, C.U.; Sundström, A.; Qi, J.; Bradner, J.E.; Weiss, W.A.; Cho, Y.-J.; Weishaupt, H.; Swartling, F.J. Combined BET bromodomain and CDK2 inhibition in MYC-driven medulloblastoma. Oncogene 2018, 37, 2850–2862. [Google Scholar] [CrossRef] [Green Version]
  71. Hoffman, L.M.; Van Zanten, S.E.V.; Colditz, N.; Baugh, J.; Chaney, B.; Hoffmann, M.; Lane, A.; Fuller, C.; Miles, L.; Hawkins, C.; et al. Clinical, Radiologic, Pathologic, and Molecular Characteristics of Long-Term Survivors of Diffuse Intrinsic Pontine Glioma (DIPG): A Collaborative Report From the International and European Society for Pediatric Oncology DIPG Registries. J. Clin. Oncol. 2018, 36, 1963–1972. [Google Scholar] [CrossRef]
  72. Schwartzentruber, J.; Korshunov, A.; Liu, X.-Y.; Jones, D.T.W.; Pfaff, E.; Jacob, K.; Sturm, D.; Fontebasso, A.M.; Khuong-Quang, D.-A.; Tönjes, M.; et al. Driver mutations in histone H3.3 and chromatin remodelling genes in paediatric glioblastoma. Nature 2012, 482, 226–231. [Google Scholar] [CrossRef]
  73. Wu, G.; Broniscer, A.; McEachron, T.A.; Lu, C.; Paugh, B.S.; Becksfort, J.; Qu, C.; Ding, L.; Huether, R.; Parker, M.; et al. Somatic histone H3 alterations in pediatric diffuse intrinsic pontine gliomas and non-brainstem glioblastomas. Nat. Genet. 2012, 44, 251–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Lewis, P.W.; Müller, M.M.; Koletsky, M.S.; Cordero, F.; Lin, S.; Banaszynski, L.A.; Garcia, B.A.; Muir, T.W.; Becher, O.J.; Allis, C.D. Inhibition of PRC2 Activity by a Gain-of-Function H3 Mutation Found in Pediatric Glioblastoma. Science 2013, 340, 857–861. [Google Scholar] [CrossRef] [Green Version]
  75. Bender, S.; Tang, Y.; Lindroth, A.M.; Hovestadt, V.; Jones, D.T.W.; Kool, M.; Zapatka, M.; Northcott, P.A.; Sturm, D.; Wang, W.; et al. Reduced H3K27me3 and DNA Hypomethylation Are Major Drivers of Gene Expression in K27M Mutant Pediatric High-Grade Gliomas. Cancer Cell 2013, 24, 660–672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Piunti, A.; Hashizume, R.; Morgan, M.A.; Bartom, E.; Horbinski, C.M.; Marshall, S.A.; Rendleman, E.J.; Ma, Q.; Takahashi, Y.-H.; Woodfin, A.R.; et al. Therapeutic targeting of polycomb and BET bromodomain proteins in diffuse intrinsic pontine gliomas. Nat. Med. 2017, 23, 493–500. [Google Scholar] [CrossRef] [PubMed]
  77. Nagaraja, S.; Vitanza, N.A.; Woo, P.J.; Taylor, K.R.; Liu, F.; Zhang, L.; Li, M.; Meng, W.; Ponnuswami, A.; Sun, W.; et al. Transcriptional Dependencies in Diffuse Intrinsic Pontine Glioma. Cancer Cell 2017, 31, 635–652.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Wiese, M.; Hamdan, F.H.; Kubiak, K.; Diederichs, C.; Gielen, G.H.; Nussbaumer, G.; Carcaboso, A.M.; Hulleman, E.; Johnsen, S.A.; Kramm, C.M. Combined treatment with CBP and BET inhibitors reverses inadvertent activation of detrimental super enhancer programs in DIPG cells. Cell Death Dis. 2020, 11, 673. [Google Scholar] [CrossRef]
  79. Chan, H.M.; La Thangue, N.B. p300/CBP proteins: HATs for transcriptional bridges and scaffolds. J. Cell Sci. 2001, 114, 2363–2373. [Google Scholar] [CrossRef]
  80. Zhang, Y.; Dong, W.; Zhu, J.; Wang, L.; Wu, X.; Shan, H. Combination of EZH2 inhibitor and BET inhibitor for treatment of diffuse intrinsic pontine glioma. Cell Biosci. 2017, 7, 56. [Google Scholar] [CrossRef]
  81. Taylor, I.C.; Hütt-Cabezas, M.; Brandt, W.D.; Kambhampati, M.; Nazarian, J.; Chang, H.T.; Warren, K.E.; Eberhart, C.G.; Raabe, E.H. Disrupting NOTCH Slows Diffuse Intrinsic Pontine Glioma Growth, Enhances Radiation Sensitivity, and Shows Combinatorial Efficacy With Bromodomain Inhibition. J. Neuropathol. Exp. Neurol. 2015, 74, 778–790. [Google Scholar] [CrossRef] [Green Version]
  82. Wang, J.; Huang, T.Y.-T.; Hou, Y.; Bartom, E.; Lu, X.; Shilatifard, A.; Yue, F.; Saratsis, A. Epigenomic landscape and 3D genome structure in pediatric high-grade glioma. Sci. Adv. 2021, 7, eabg4126. [Google Scholar] [CrossRef]
  83. Elsamadicy, A.A.; Koo, A.B.; David, W.B.; Lee, V.; Zogg, C.K.; Kundishora, A.J.; Hong, C.S.; DeSpenza, T.; Reeves, B.; Kahle, K.T.; et al. Comparison of epidemiology, treatments, and outcomes in pediatric versus adult ependymoma. Neuro Oncol. Adv. 2020, 2, vdaa019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Tihan, T.; Zhou, T.; Holmes, E.; Burger, P.C.; Ozuysal, S.; Rushing, E.J. The prognostic value of histological grading of posterior fossa ependymomas in children: A Children’s Oncology Group study and a review of prognostic factors. Mod. Pathol. 2007, 21, 165–177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Vitanza, N.A.; Partap, S. Pediatric Ependymoma. J. Child Neurol. 2016, 31, 1354–1366. [Google Scholar] [CrossRef]
  86. National Library of Medicine (U.S.). Maintenance Chemotherapy or Observation Following Induction Chemotherapy and Radiation Therapy in Treating Patients with Newly Diagnosed Ependymoma. March 2010. Identifier NCT01096368. Available online: https://clinicaltrials.gov/ct2/show/NCT01096368 (accessed on 10 March 2022).
  87. Pajtler, K.W.; Witt, H.; Sill, M.; Jones, D.T.; Hovestadt, V.; Kratochwil, F.; Wani, K.; Tatevossian, R.; Punchihewa, C.; Johann, P.; et al. Molecular Classification of Ependymal Tumors across All CNS Compartments, Histopathological Grades, and Age Groups. Cancer Cell 2015, 27, 728–743. [Google Scholar] [CrossRef] [Green Version]
  88. Arabzade, A.; Zhao, Y.; Varadharajan, S.; Chen, H.-C.; Jessa, S.; Rivas, B.; Stuckert, A.J.; Solis, M.; Kardian, A.; Tlais, D.; et al. ZFTA–RELA Dictates Oncogenic Transcriptional Programs to Drive Aggressive Supratentorial Ependymoma. Cancer Discov. 2021, 11, 2200–2215. [Google Scholar] [CrossRef] [PubMed]
  89. Kupp, R.; Ruff, L.; Terranova, S.; Nathan, E.; Ballereau, S.; Stark, R.; Chilamakuri, C.S.R.; Hoffmann, N.; Wickham-Rahrmann, K.; Widdess, M.; et al. ZFTA Translocations Constitute Ependymoma Chromatin Remodeling and Transcription Factors. Cancer Discov. 2021, 11, 2216–2229. [Google Scholar] [CrossRef]
  90. Zheng, T.; Ghasemi, D.R.; Okonechnikov, K.; Korshunov, A.; Sill, M.; Maass, K.K.; da Silva, P.B.G.; Ryzhova, M.; Gojo, J.; Stichel, D.; et al. Cross-Species Genomics Reveals Oncogenic Dependencies in ZFTA/C11orf95 Fusion–Positive Supratentorial Ependymomas. Cancer Discov. 2021, 11, 2230–2247. [Google Scholar] [CrossRef]
  91. Bockmayr, M.; Harnisch, K.; Pohl, L.C.; Schweizer, L.; Mohme, T.; Körner, M.; Alawi, M.; Suwala, A.K.; Dorostkar, M.M.; Monoranu, C.M.; et al. Comprehensive profiling of myxopapillary ependymomas identifies a distinct molecular subtype with relapsing disease. Neuro Oncol. 2022. Available online: https://pubmed.ncbi.nlm.nih.gov/35380708/ (accessed on 10 March 2022). [CrossRef]
  92. Mack, S.C.; Pajtler, K.W.; Chavez, L.; Okonechnikov, K.; Bertrand, K.C.; Wang, X.; Erkek, S.; Federation, A.; Song, A.; Lee, C.; et al. Therapeutic targeting of ependymoma as informed by oncogenic enhancer profiling. Nature 2017, 553, 101–105. [Google Scholar] [CrossRef]
  93. Servidei, T.; Meco, D.; Martini, M.; Battaglia, A.; Granitto, A.; Buzzonetti, A.; Babini, G.; Massimi, L.; Tamburrini, G.; Scambia, G.; et al. The BET inhibitor OTX015 exhibits in vitro and in vivo antitumor activity in pediatric ependymoma stem cell models. Int. J. Mol. Sci. 2021, 22, 1877. [Google Scholar] [CrossRef]
  94. Eberhart, C.G.; Brat, D.J.; Cohen, K.J.; Burger, P.C. Pediatric Neuroblastic Brain Tumors Containing Abundant Neuropil and True Rosettes. Pediatr. Dev. Pathol. 2000, 3, 346–352. [Google Scholar] [CrossRef]
  95. Li, M.; Lee, K.F.; Lu, Y.; Clarke, I.; Shih, D.J.H.; Eberhart, C.; Collins, V.P.; Van Meter, T.; Picard, D.; Zhou, L.; et al. Frequent Amplification of a chr19q13.41 MicroRNA Polycistron in Aggressive Primitive Neuroectodermal Brain Tumors. Cancer Cell 2009, 16, 533–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Pfister, S.; Remke, M.; Castoldi, M.; Bai, A.H.C.; Muckenthaler, M.; Kulozik, A.; von Deimling, A.; Pscherer, A.; Lichter, P.; Korshunov, A. Novel genomic amplification targeting the microRNA cluster at 19q13.42 in a pediatric embryonal tumor with abundant neuropil and true rosettes. Acta Neuropathol. 2008, 117, 457–464. [Google Scholar] [CrossRef] [PubMed]
  97. Lambo, S.; Gröbner, S.N.; Rausch, T.; Waszak, S.; Schmidt, C.; Gorthi, A.; Romero, J.C.; Mauermann, M.; Brabetz, S.; Krausert, S.; et al. The molecular landscape of ETMR at diagnosis and relapse. Nature 2019, 576, 274–280. [Google Scholar] [CrossRef] [PubMed]
  98. Horwitz, M.; Dufour, C.; Leblond, P.; Bourdeaut, F.; Faure-Conter, C.; Bertozzi, A.-I.; Delisle, M.B.; Palenzuela, G.; Jouvet, A.; Scavarda, D.; et al. Embryonal tumors with multilayered rosettes in children: The SFCE experience. Child’s Nerv. Syst. 2015, 32, 299–305. [Google Scholar] [CrossRef] [Green Version]
  99. Hanson, D.; Hoffman, L.M.; Nagabushan, S.; Goumnerova, L.C.; Rathmann, A.; Vogel, T.; Ziegler, D.S.; Chi, S. A modified IRS-III chemotherapy regimen leads to prolonged survival in children with embryonal tumor with multilayer rosettes. Neuro Oncol. Adv. 2020, 2, vdaa120. [Google Scholar] [CrossRef]
  100. Sin-Chan, P.; Mumal, I.; Suwal, T.; Ho, B.; Infante, E.R.; Singh, I.; Du, Y.; Lu, M.; Patel, N.; Torchia, J.; et al. A C19MC-LIN28A-MYCN Oncogenic Circuit Driven by Hijacked Super-enhancers Is a Distinct Therapeutic Vulnerability in ETMRs: A Lethal Brain Tumor. Cancer Cell 2019, 36, 51–67.e7. [Google Scholar] [CrossRef] [Green Version]
  101. Viswanathan, S.; Powers, J.T.; Einhorn, W.; Hoshida, Y.; Ng, T.L.; Toffanin, S.; O’Sullivan, M.; Lu, J.; Phillips, L.A.; Lockhart, V.L.; et al. Lin28 promotes transformation and is associated with advanced human malignancies. Nat. Genet. 2009, 41, 843–848. [Google Scholar] [CrossRef] [Green Version]
  102. Ginn, K.F.M.; Gajjar, A.M. Atypical Teratoid Rhabdoid Tumor: Current Therapy and Future Directions. Front. Oncol. 2012, 2, 114. [Google Scholar] [CrossRef] [Green Version]
  103. Biegel, J.A.; Tan, L.; Zhang, F.; Wainwright, L.; Russo, P.; Rorke, L.B. Alterations of the hSNF5/INI1 gene in central nervous system atypical teratoid/rhabdoid tumors and renal and extrarenal rhabdoid tumors. Clin. Cancer Res. 2002, 8, 3461–3467. [Google Scholar]
  104. Versteege, I.; Sévenet, N.; Lange, J.; Rousseau-Merck, M.-F.; Ambros, P.; Handgretinger, R.; Aurias, A.; Delattre, O. Truncating mutations of hSNF5/INI1 in aggressive paediatric cancer. Nature 1998, 394, 203–206. [Google Scholar] [CrossRef]
  105. Lee, R.S.; Stewart, C.; Carter, S.L.; Ambrogio, L.; Cibulskis, K.; Sougnez, C.; Lawrence, M.S.; Auclair, D.; Mora, J.; Golub, T.R.; et al. A remarkably simple genome underlies highly malignant pediatric rhabdoid cancers. J. Clin. Investig. 2012, 122, 2983–2988. [Google Scholar] [CrossRef] [PubMed]
  106. Johann, P.D.; Erkek, S.; Zapatka, M.; Kerl, K.; Buchhalter, I.; Hovestadt, V.; Jones, D.T.; Sturm, D.; Hermann, C.; Wang, M.S.; et al. Atypical Teratoid/Rhabdoid Tumors Are Comprised of Three Epigenetic Subgroups with Distinct Enhancer Landscapes. Cancer Cell 2016, 29, 379–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Torchia, J.; Picard, D.; Lafay-Cousin, L.; Hawkins, C.E.; Kim, S.-K.; Letourneau, L.; Ra, Y.-S.; Ho, K.C.; Chan, T.S.Y.; Sin-Chan, P.; et al. Molecular subgroups of atypical teratoid rhabdoid tumours in children: An integrated genomic and clinicopathological analysis. Lancet Oncol. 2015, 16, 569–582. [Google Scholar] [CrossRef]
  108. Torchia, J.; Golbourn, B.; Feng, S.; Ho, K.C.; Sin-Chan, P.; Vasiljevic, A.; Norman, J.D.; Guilhamon, P.; Garzia, L.; Agamez, N.R.; et al. Integrated (epi)-Genomic Analyses Identify Subgroup-Specific Therapeutic Targets in CNS Rhabdoid Tumors. Cancer Cell 2016, 30, 891–908. [Google Scholar] [CrossRef] [Green Version]
  109. Alimova, I.; Pierce, A.; Danis, E.; Donson, A.; Birks, D.K.; Griesinger, A.; Foreman, N.K.; Santi, M.; Soucek, L.; Venkataraman, S.; et al. Inhibition of MYC attenuates tumor cell self-renewal and promotes senescence in SMARCB1-deficient Group 2 atypical teratoid rhabdoid tumors to suppress tumor growth in vivo. Int. J. Cancer 2019, 144, 1983–1995. [Google Scholar] [CrossRef]
  110. Berenguer-Daizé, C.; Astorgues-Xerri, L.; Odore, E.; Cayol, M.; Cvitkovic, E.; Noel, K.; Bekradda, M.; MacKenzie, S.; Rezai, K.; Lokiec, F.; et al. OTX015 (MK-8628), a novel BET inhibitor, displays in vitro and in vivo antitumor effects alone and in combination with conventional therapies in glioblastoma models. Int. J. Cancer 2016, 139, 2047–2055. [Google Scholar] [CrossRef]
  111. Hottinger, A.F.; Sanson, M.; Moyal, E.; Delord, J.-P.; De Micheli, R.; Rezai, K.; Leung, A.C.; Perez, S.; Bekradda, M.; Lachaux, N.; et al. Dose optimization of MK-8628 (OTX015), a small molecule inhibitor of bromodomain and extra-terminal (BET) proteins, in patients (pts) with recurrent glioblastoma (GB). J. Clin. Oncol. 2016, 34, e14123. [Google Scholar] [CrossRef]
  112. Moreno, V.; Sepulveda, J.; Vieito, M.; Hernández-Guerrero, T.; Doger, B.; Saavedra, O.; Ferrero, O.; Sarmiento, R.; Arias, M.; De Alvaro, J.; et al. Phase I study of CC-90010, a reversible, oral BET inhibitor in patients with advanced solid tumors and relapsed/refractory non-Hodgkin’s lymphoma. Ann. Oncol. 2020, 31, 780–788. [Google Scholar] [CrossRef]
  113. Pearson, A.D.; DuBois, S.G.; Buenger, V.; Kieran, M.; Stegmaier, K.; Bandopadhayay, P.; Bennett, K.; Bourdeaut, F.; Brown, P.A.; Chesler, L.; et al. Bromodomain and extra-terminal inhibitors—A consensus prioritisation after the Paediatric Strategy Forum for medicinal product development of epigenetic modifiers in children—ACCELERATE. Eur. J. Cancer 2021, 146, 115–124. [Google Scholar] [CrossRef]
Table 1. BET inhibitors in clinical development.
Table 1. BET inhibitors in clinical development.
AgentSelectivityPhaseDisease Focus; CommentsStructure
ABBV-075Pan-BETISolid tumors, AML, multiple myeloma, myelofibrosis Pharmaceuticals 15 00665 i001
ABBV-744BDII selectiveIAML, myelofibrosis Pharmaceuticals 15 00665 i002
AZDZ5153Pan-BETI/IIAML; Unique bivalent binding mode Pharmaceuticals 15 00665 i003
BI-894999Pan-BETIAdvanced solid tumors, DLBCL, or NMC Pharmaceuticals 15 00665 i004
BMS-986158Pan-BETI/IIAdvanced solid tumors and hematologic malignancies; Ongoing pediatric study Pharmaceuticals 15 00665 i005
BMS-986378 (CC-90010)Pan-BETIAdvanced solid tumors, NHL; Ongoing pediatric study, good CNS penetration Pharmaceuticals 15 00665 i006
CPI-0610Pan-BETI/IILymphoma, Multiple myeloma, AML, MPNST Pharmaceuticals 15 00665 i007
I-BET762 (GSK525762)Pan-BETI/IIHematologic malignancies, solid tumors Pharmaceuticals 15 00665 i008
INCB57643Pan-BETI/IIHematologic malignancies, solid tumors Pharmaceuticals 15 00665 i009
NEO2734Pan-BET and P300I/IIHematologic malignancies, solid tumors Pharmaceuticals 15 00665 i010
OTX015Pan-BETI/IIHematologic malignancies, solid tumors, GBM Pharmaceuticals 15 00665 i011
PLX51107Pan-BETIHematologic malignancies, solid tumors Pharmaceuticals 15 00665 i012
Source: Clinicaltrials.gov (accessed on 10 March 2022); AML = acute myeloid leukemia, DLBCL = diffuse large B-cell lymphoma, GBM = glioblastoma multiforme, NMC = NUT midline carcinoma, MPNST = Malignant peripheral nerve sheath tumor.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Groves, A.; Clymer, J.; Filbin, M.G. Bromodomain and Extra-Terminal Protein Inhibitors: Biologic Insights and Therapeutic Potential in Pediatric Brain Tumors. Pharmaceuticals 2022, 15, 665. https://doi.org/10.3390/ph15060665

AMA Style

Groves A, Clymer J, Filbin MG. Bromodomain and Extra-Terminal Protein Inhibitors: Biologic Insights and Therapeutic Potential in Pediatric Brain Tumors. Pharmaceuticals. 2022; 15(6):665. https://doi.org/10.3390/ph15060665

Chicago/Turabian Style

Groves, Andrew, Jessica Clymer, and Mariella G. Filbin. 2022. "Bromodomain and Extra-Terminal Protein Inhibitors: Biologic Insights and Therapeutic Potential in Pediatric Brain Tumors" Pharmaceuticals 15, no. 6: 665. https://doi.org/10.3390/ph15060665

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

Article Metrics

Back to TopTop