Opinion Review Open Access
Copyright ©The Author(s) 2021. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Stem Cells. Jul 26, 2021; 13(7): 670-684
Published online Jul 26, 2021. doi: 10.4252/wjsc.v13.i7.670
Epigenetic modulators for brain cancer stem cells: Implications for anticancer treatment
Luana Abballe, Evelina Miele, Department of Pediatric Hematology/Oncology and Cellular and Gene Therapy, Bambino Gesù Children's Hospital, IRCCS, Rome 00165, Italy
ORCID number: Luana Abballe (0000-0003-4446-2293); Evelina Miele (0000-0002-4747-1032).
Author contributions: Abballe L collected the data and wrote the paper; Miele E contributed to the conception and design and was responsible for review supervision; both authors read and approved the final version of the manuscript.
Supported by Italian Ministry of Health, Ricerca Finalizzata, No. GR-2018-12367328 (to Miele E).
Conflict-of-interest statement: Authors declare no conflict of interests for this article.
Open-Access: This article is an open-access article that was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution NonCommercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/Licenses/by-nc/4.0/
Corresponding author: Evelina Miele, MD, PhD, Medical Assistant, Postdoc, Research Scientist, Department of Pediatric Hematology/Oncology and Cellular and Gene Therapy, Bambino Gesù Children's Hospital, IRCCS, Piazza di Sant'Onofrio, 4, Rome 00165, Italy. evelina.miele@opbg.net
Received: March 13, 2021
Peer-review started: March 13, 2021
First decision: April 6, 2021
Revised: April 26, 2021
Accepted: June 22, 2021
Article in press: June 22, 2021
Published online: July 26, 2021

Abstract

Primary malignant brain tumors are a major cause of morbidity and mortality in both adults and children, with a dismal prognosis despite multimodal therapeutic approaches. In the last years, a specific subpopulation of cells within the tumor bulk, named cancer stem cells (CSCs) or tumor-initiating cells, have been identified in brain tumors as responsible for cancer growth and disease progression. Stemness features of tumor cells strongly affect treatment response, leading to the escape from conventional therapeutic approaches and subsequently causing tumor relapse. Recent research efforts have focused at identifying new therapeutic strategies capable of specifically targeting CSCs in cancers by taking into consideration their complex nature. Aberrant epigenetic machinery plays a key role in the genesis and progression of brain tumors as well as inducing CSC reprogramming and preserving CSC characteristics. Thus, reverting the cancer epigenome can be considered a promising therapeutic strategy. Three main epigenetic mechanisms have been described: DNA methylation, histone modifications, and non-coding RNA, particularly microRNAs. Each of these mechanisms has been proven to be targetable by chemical compounds, known as epigenetic-based drugs or epidrugs, that specifically target epigenetic marks. We review here recent advances in the study of epigenetic modulators promoting and sustaining brain tumor stem-like cells. We focus on their potential role in cancer therapy.

Key Words: Cancer stem cells, Epigenetics, Brain tumors, Epigenetic drugs, Histone deacetylase inhibitors, DNA methyltransferase inhibitors

Core Tip: Cancer stem cells (CSCs) are characterized by an altered epigenome that contributes to treatment failure and tumor relapse. Physicians are looking for new therapeutic approaches to target specifically CSCs in cancers. In this review, we summarize literature data about epigenetic markers in brain CSCs and shed light on new epigenetic therapies.



INTRODUCTION

Primary malignant brain tumors are a heterogeneous group of tumors arising from the brain parenchyma and its surrounding structures. They represent a major cause of morbidity and mortality in both adults and children. In particular, brain tumors are the most frequently reported solid malignancies in children, accounting for up to 20% of childhood cancers[1]. Brain cancers require a multimodal treatment that includes surgical resection, chemotherapy, and radiation therapy[2]. This therapeutic strategy includes a particularly invasive surgery as well as brain irradiation and can lead to long-lasting side effects, significantly decreasing the patient’s quality of life.

Brain tumors can harbor genetic and epigenetic alterations that make them resistant to conventional pharmacological treatment[3]. Current therapeutic approaches do not consider the presence of a specific subpopulation of cells within the tumor bulk, known as cancer stem cells (CSCs) or tumor-initiating cells. These cells are not responsive to conventional treatment, promoting tumor relapse[4]. In this scenario, it is of fundamental importance to seek alternative therapeutic strategies that take into account the genetic and epigenetic alterations of the tumor as well as the presence of tumor-initiating cells. Nowadays, to address this challenge, the epigenetic profile of CSCs is being thoroughly studied and several epigenetics drugs are being tested in in vitro and in vivo studies on CSC cultures[5].

This review aims to summarize and organize the current epigenetics strategies for eradicating brain CSCs based on CSC epigenetic modulators.

Brain cancer stem cells

CSCs are considered the “reservoirs” of the tumor. They are defined as a small subset of stem cells capable of proliferating and generating diverse and heterogeneous cell types that constitute the whole tumor[6]. CSCs reside in specific anatomical areas called “niches”, where they interact with the microenvironment surrounding them[7]. Like their normal tissue counterpart, CSCs exhibit “stem-like” characteristics: (1) Cell quiescence: A way to preserve self-renewal and to avoid genetic perturbation that could occur during cell division; (2) Self-renewal capacity: The ability to proliferate symmetrically and asymmetrically; (3) Multipotency: The ability to give rise to heterogeneous cells with different proliferative potential; (4) Migration: The ability to migrate and disseminate; (5) Tissue regeneration: The ability to give rise to new tumoral tissues; and (6) Communication: The ability to interact with the microenvironment.

The substantial difference between CSCs and normal tissue stem cells is that proliferation/death signals are aberrant and dysregulated in cancer, whereas normal tissue stem cells can maintain physiological homeostasis.

CSCs were discovered for the first time in acute myeloid leukemia[8]. Subsequently, this tumor-initiation subpopulation was also identified in many solid tumors, including brain cancers[9,10]. CSCs have been discovered and isolated in major brain tumors [e.g., medulloblastoma (MB), gliomas, ependymoma][11-13] from both pediatric and adult patients[14-16].

In the first instance, cell surface markers were used to identify this subset of tumor-initiating cells. The cell surface antigen CD133 has been most frequently used to mark CSCs in various solid tumors[17]. In the context of the brain, Singh et al[11] demonstrated that CD133+ human brain tumor cells were able to initiate and recapitulate the original tumor in in vivo mouse models. Further study demonstrated that a high expression level of CD133 was correlated with poor prognosis in brain cancer patients, reinforcing CD133’s role as a “brain stemness” marker[18].

The accuracy of this detection method remains very controversial because cell surface markers evolve rapidly in response to different environmental stimuli, disease states, and tumor progression[19,20]. Moreover, brain CSCs and neural stem cells share common phenotypic markers (e.g., CD133, CD15, CD44)[21]. Hence, the marker-dependent identification method alone is insufficient to discriminate correctly CSCs from normal tissue ones. Recent studies[22-24] investigate complementary methods that consider cell dysregulated survival pathways and genetic and epigenetic signatures. As already described by Abbaszadegan et al[23], the gold standard strategy to identify efficiently brain CSCs is to test their in vivo tumorigenicity. Limiting dilution assay is the best tumorigenicity method that is commonly used for evaluation of CSCs frequency. However, this method presents some critical points, being influenced by the number of the cells, the implantation site, and growing time of incubation. Moreover, it is not feasible on large scale studies. Complementary in vitro functional assays could be used to identify CSCs based on: (1) Their intrinsic properties (e.g., self-renewal, asymmetrical division, slow proliferation phenotype, and aldehyde dehydrogenase 1 expression); and (2) Their survival pathways (e.g., Wnt/β-catenin, Hedgehog, and Notch signaling pathway), in terms of expression of transcription factors/key proteins/microRNAs (miRNAs). Among the recently developed approaches to isolate CSCs, there are next generation sequencing (NGS) technologies. For example, Jonasson et al[24] isolated the stem-like subpopulation using a functional cellular assay that enriches for cells that can self-renew and differentiate, combined with NGS technologies (single-cell RNA sequencing) to identify CSCs[22-24]. Moreover, Rodriguez-Meira et al[25] in their scientific work developed an NGS platform that combines single-cell RNA-seq with mutational analysis allowing the identification of distinct subclones of cancer cells[26]. This evidence suggests that a combination of cell surface markers and functional assays provide an efficient tool for their identification.

CSCs’ properties have profound implications for treatment response due to their ability to evade conventional therapy and cause subsequent tumor relapse[27,28]. In this case, it is crucial to identify new treatment strategies that take into consideration the complexities of CSCs. Currently, global researchers are making great efforts to understand the biological properties of CSCs and develop new therapeutic approaches targeting CSCs.

EPIGENETIC MODULATORS

The term "epigenetics" is used to refer to information that controls gene expression that is stable and inheritable during cell division and happens without changes in the DNA sequence. Aberrant epigenetic landscapes control cell fate specification, promoting tumor initiation and progression[29,30].

Epigenetics controls gene expression through three main mechanisms: DNA methylation, histone modifications, and non-coding RNA, particularly miRNAs[31] (Figure 1).

Figure 1
Figure 1 Epigenetic mechanism in cancer stem cells and therapeutical approaches. Epigenetic modifiers: Non-coding RNA, DNA methylation, and histone modifications. Non-coding RNAs (microRNAs) regulate gene expression. microRNA expression can be regulated by other epigenetic modifiers. DNA methyltransferases add a methyl group to cytosine in the DNA sequence. This results in aberrant gene silencing and changes in normal gene expression. Posttranslational modifications of histone proteins (acetylation in purple and methylation in red) can affect chromatin structure. Histone enzymes add/remove acetyl or methyl groups. Histone deacetylases and DNMTs are the main targets of epigenetic inhibitors. CSC: Cancer stem cell; Me: Methyl group; Ac: Acetyl group; DNMT: DNA methyltransferases; HDAC: Histone deacetylases; HAT: Histone acetyltransferases; HMT: Histone methyltransferases; HDM: Histone demethylases; DNMTi: DNA methyltransferases inhibitors; HDACi: Histone deacetylases inhibitors.

DNA methylation is the most studied epigenetic modification. It is associated with transcriptional inactivation and closed chromatin structure, mainly regulating gene silencing or repression. It depends on the action of specialized enzymes, known as DNA methyltransferases (DNMTs) that transfer a methyl group from S-adenosyl methionine to the fifth carbon of the pyrimidine ring of a DNA base cytosine. DNA methylation occurs mainly at CpG dinucleotides, which are concentrated in specific regions of DNA called “CpG islands”. CpG islands are located at gene promoters, in regulatory regions, and in gene bodies[32], but DNA methylation could also be present in non-coding DNA sequences such as repetitive elements, transposons, non-coding RNAs, and introns[33,34].

A methylation pattern represents the fingerprint of a cell. It is established during early embryogenesis, and it is maintained during cell division by the action of DNMTs.

Methylation status is altered in cancers, mainly through two mechanisms: Regional hypermethylation and global hypomethylation[35]. DNA hypermethylation involves the CpG islands, which are usually unmethylated in normal cells. The result of DNA hypermethylation is the transcriptional repression of tumor-suppressor and tissue-specific genes, and the inactivation of miRNA, which are involved in the initiation and progression of cancer[34]. This mechanism occurs in different stages of carcinogenesis including CSC formation[36]. Conversely, global hypomethylation consists of the loss of the methyl group on cytosine, mainly in repetitive elements across DNA. Hypomethylation was one of the first epigenetic features discovered in human cancers[37], causing the reactivation of methylated regions of DNA such as transposons, introns, and germ-line genes that are silent in differentiated cells[34,38]. In high-grade cancers, such as glioblastoma (GBM), hypomethylation is a mechanism used by CSCs to reactivate key stem cell genes[39].

Abnormalities of DNA methylation are early events in pre-malignant transformation and are maintained in the global tumor population. However, the epigenome is in continuous evolution, and some of the changes are detectable in later steps of tumorigenesis as a result of positive selection. In this way, epigenome contributes to tumor heterogeneity and plasticity, which gives rise to a heterogeneous tumor composed of different cell subpopulations, one of them could have “stem-like” features. Additionally, compared to the bulk tumor, CSCs could acquire further epigenetic alterations in response to stress of different nature (e.g., chemotherapy/radiotherapy, chronic inflammation, and environmental exposures), contributing to tumor relapse[40].

The development of powerful “next-generation” techniques makes it possible to describe the entire epigenome of cells or tissues[38]. Capper et al[41] developed a DNA methylation-based classification for central nervous system cancers that allows discrimination between different subtypes of tumors, some of them previously considered as homogeneous diseases.

Histones are subject to reversible post-translational modifications (PTMs) that cooperate to govern the chromatin state. Histone PTMs influence chromatin structure that is conducive to the expression or repression of target genes. Histone amino-terminal regions can undergo diverse PTMs: Methylation, acetylation, phosphorylation, ubiquitylation, biotinylation, sumoylation, and ADP-ribosylation[42] that work in concert to define the chromatin status of the specific region of DNA.

Histone PTMs are controlled by different enzymes: "Writers" catalyze histone modifications; “erasers” cut histone modifications; and “readers” translate the PTMs’ language into cellular signals. All enzymes work together in a very specific manner to create the “histone code”.

Histone methylation and acetylation are the best-known PTMs on histone residues. Histone methylation is a reversible mechanism that occurs on lysine and arginine residues, leading to a different degree of methylation. Up to three methyl groups can be added to a single lysine residue (un-, mono-, di-, and tri-methylated states) and up to two groups to a single arginine residue (mono- and di-methylated states). Methylation can be either an activating or a repressive mark, depending on the target histone and lysine residue: For example, histone H3 Lysine 27 (H3K27) and histone H4 Lysine 20 (H4K20) are usually associated with gene silencing, while H3 Lysine 4 (H3K4) and H3 Lysine 36 (H3K36) are transcriptional activation marks. Histone acetylation is associated with relaxed chromatin structure and promotes the binding of transcription factors and RNA polymerase to DNA, resulting in the activation of gene expression[42].

Recent efforts have tried to highlight the role of epigenetic modulators in the development and plasticity of brain tumor CSCs[43,44]. In GBM CSCs (GSCs), Liau et al[45] showed in patient-derived GSCs, which effectively initiate tumors in mice models, that KDM6-mediated demethylation of histone H3 Lysine 27 trimethylation (H3K27me3), a repressive mark, has a central role in the maintenance and persistence of GSCs. Marampon et al[46] demonstrated the fundamental role of two histone deacetylases (HDAC4 and HDAC6) in regulating the DNA repair machinery, survival, and stemness characteristics of radioresistant GSCs derived from U87MG and U251MG human GMB cell lines subjected to irradiation. Moreover, a recent study by Banelli et al[47] on human GSCs derived from human tumors describes a subset of GSCs resistant to chemotherapy that are particularly dependent on KDM5A action, making them strongly sensitive to HDAC inhibitor (HDACi) treatments in term of cell viability, percentage of apoptotic cells, and reducing capacity of clonal growth. In medulloblastoma, pharmacological inhibition of histone methyltransferase enhancer of zeste 2 (EZH2) impairs proliferation and self-renewal of human and mouse stem-like cells, derived from primary human Sonic Hedgehog MB (SHH-MBs) and from tumors arisen in Ptc+/- mice, in in vitro and in vivo studies[48].

miRNAs and epigenetic modulators create a miRNA-epigenetic feedback loop: On the one hand, miRNA can regulate the transcription of epigenetics-associated enzymes, while on the other miRNA transcription is under the control of epigenetics machinery (DNA methylation, histone PTMs, and RNA modifications)[49].

In literature, there are several examples of miRNA-epigenetic machinery in CSCs. For example, Ferretti et al[50] described the role of the miR-326-ARRB1-E2F1 axis in medulloblastoma CSCs survival. For this study, MB CSCs were derived from tissues freshly resected from pediatric patients and cultured in CSC-enriched cultures. Notably, miR-326 is already described as onco-suppressor miRNA[50]. It is involved in neuronal differentiation, and its expression is under control of the histone-lysine N-methyltransferase EZH2. High levels of EZH2 are responsible for the presence of H3K27me3 on the promoter region of miR-326[22]. In another study, Lizarte Neto et al[51] investigated the role of miR-181d and the methylation status of the MGMT gene in response to pharmacological treatment in CD133+ GBM CSCs. Their results showed an increase of miR-181d and MGMT transcription levels after treatment, due to cellular mechanism of drug resistance[51].

The tumor microenvironment (TME) also acts as an epigenetic regulator for cancer cells. TME communicates with cancer cells through extracellular vesicles secreted by many TME cell types that contain various mediators including proteins and nucleic acids. Also, miRNAs can be charged in the extracellular vesicles and thereby alter the epigenome of the recipient cancer cell[52].

All this evidence makes brain CSCs an excellent candidate for pharmacological epigenetic treatments.

TRANSLATIONAL SIGNIFICANCE OF EPIGENETICS: EPIDRUGS

Epigenetics-based drugs (epidrugs) are chemical compounds that specifically target epigenetic marks (Figure 1). CSCs could acquire mutations in epigenetic marks or changing in methylation/acetylation status, or again in miRNAs signature, that make them sensitive to epigenetics-based drugs’ approaches. Epidrugs are based on the idea that by changing the epigenome of CSCs, it may be possible to remodel the fate of cells from CSCs to differentiated tumor cells[53-55]. Epidrugs are used as monotherapy or in combination with chemo-/radio-therapy to target both CSCs and bulk tumors. Today, several HDAC and DNMT inhibitors are in different clinical trial phases.

HDACi

Among the various epigenetic modulators, the acetylation state of histone proteins is one of the major targets for anticancer therapy. Histone acetylation is controlled by two types of enzymes: HDACs and histone acetyltransferases (Figure 1).

Human HDACs are categorized into four different classes: Class I (HDAC1, 2, 3, and 8), class IIa (HDAC4, 5, 7, and 9), class IIb (HDAC6 and 10), class III (sirt 1-7), and class IV (HDAC11). HDACs are enzymes that remove the acetyl group from lysine residues on histones, which compact the chromatin structure into a non-permissive state, resulting in transcriptional repression.

Based on isoform selectivity, HDACi can be classified as: (1) Pan-inhibitors, if they act against all HDACs; or (2) HDAC isoform-selective inhibitors if they target a specific HDAC class[56]. Pan-inhibitors are in turn classified according to their chemical structure as: (1) Hydroxamic acids; (2) Aliphatic carboxylic acids; (3) Benzamides; (4) Cyclic peptides; or (5) Sirtuin inhibitors.

HDACs control several cellular mechanisms. They have been implicated in different types of cancers[46], and several HDACs are overexpressed in brain cancers[57,58]. For example, Staberg et al[59] found an up-expression of HDAC1, 3, and 6 in 21 primary GBM cell cultures, grown as neurospheres, compared to non-neoplastic brain control cells (normal human astrocytes), and confirmed these findings in a panel of primary GBM tissue samples compared to normal brain tissues. They demonstrated the efficacy of HDACi therapy (trichostatin A) in GBM treatment in in vitro experiments.

HDACi exhibit anti-cancer activity against CSCs in tumors with a predominant stem-like population such as brain cancers. HDACi target the escape mechanism of CSCs, reversing chemo-radio-therapy resistance by inducing cell differentiation, apoptosis, inhibition of angiogenesis, and upregulation of tumor suppressor genes[60].

In line with this observation, da Cunha Jaeger et al[61] demonstrated that HDACi and mitogen-activated protein kinase/extracellular signal-regulated kinase inhibitors could modulate the stemness markers (BMI1 and CD133), viability, and neurosphere formation capacity of MB cell lines (DAOY and D283 cultured in serum-free sphere-induction medium). Their results proposed HDACi as a valid candidate for MB CSC treatment[61]. Coni et al[62] reported antitumor effects of selective HDAC1 and HDAC2 inhibition on SHH-MB cells and mouse models. Also, in diffuse intrinsic pontine glioma tumor models (patient-derived neurospheres, xenografts, and allografts), combination of HDAC and AXL inhibition modulate the H3K27M epigenetic mark resulting in a down-regulation of stemness markers (SOX2 and its target genes directly correlated with stem cell characteristics)[63].

Several HDACi are emerging as promising anticancer drugs in clinical trials for brain cancer treatment, both as combinatorial- and as mono-therapy[64]. Table 1 shows the HDACi used in clinical trials for brain cancers (Vorinostat, Valproic acid, Panobinostat, and Entinostat). Most clinical trials are focused on the usage of Vorinostat (suberoylanilide hydroxamic acid or SAHA), a potent inhibitor of HDAC classes I and II. Vorinostat was initially approved by the Food and Drug Administration for treating refractory cutaneous T-cell lymphoma, and its use was subsequently extended to various solid cancers[56]. In a phase 2 clinical trial, Vorinostat was tested as a mono-therapeutic agent in GBM treatment to target both CSCs and differentiated cancer cells. In fact, preclinical studies demonstrated the efficacy of Vorinostat in reducing EZH2 and CD133 stemness marker in patient-derived GBM CSCs[65] and in modulating senescence via the p38-p53 axis and inducing apoptosis in GBM cell lines cultured in serum-free sphere-induction medium[66]. Sung et al[67] demonstrated that a combination of Vorinostat with melatonin can overcome the pharmacological resistance of human GBM CSCs, reducing self-renewal and proliferation of stem cell compartments and increasing apoptosis markers (cleaved poly-ADP ribose polymerase and p-γH2AX).

Table 1 Histone deacetylases inhibitors in clinical trial for brain cancer treatment.
Compound
Class
Target
Conditions
Phase of clinical trial
Study title
EntinostatiHDAC (benzamide derivates)Class IBrain stem neoplasmPhase IEntinostat in Treating Pediatric Patients With Recurrent or Refractory Solid Tumors
Pineal region neoplasm
Recurrent lymphoma
Recurrent malignant solid neoplasm
Recurrent primary central nervous system neoplasm
Recurrent visual pathway glioma
Refractory lymphoma
Refractory malignant solid neoplasm
Refractory primary central nervous system neoplasm
Refractory visual pathway glioma
Panobinostat (LBH589)iHDAC (hydroxymic acids)Class I, II, IVRecurrent gliomaPhase IPanobinostat and Stereotactic Radiation Therapy in Treating Patients With Brain Tumors
High-grade meningioma
Brain metastasis
Panobinostat (LBH589)iHDAC (hydroxymic acids)Class I, II, IVDiffuse intrinsic pontine gliomaPhase IPhase I Study of Marizomib + Panobinostat for Children With DIPG
Pediatric brainstem glioma
Pediatric brainstem gliosarcoma recurrent pediatric cancer
Pediatric brain tumor
Diffuse glioma
Panobinostat (LBH589)iHDAC (hydroxymic acids)Class I, II, IVRecurrent malignant gliomasPhase IIPhase II Trial of LBH589 (Panobinostat) in Adult Patients With Recurrent Malignant Gliomas
Valproic AcidiHDAC (fatty acid derivates)Class I, IIGBM WHO grade IVPhase IIIInternational Cooperative Phase III Trial of the HIT-HGG Study Group (HIT-HGG-2013)
Diffuse midline glioma histone 3 K27M WHO grade IV
Anaplastic astrocytoma WHO grade III
Diffuse intrinsic pontine glioma
Gliomatosis cerebri
Valproic AcidiHDAC (fatty acid derivates)Class I, IIBrain metastasisPhase IPhase I Study of Temozolomide, Valproic Acid and Radiation Therapy in Patients With Brain Metastases
Valproic AcidiHDAC (fatty acid derivates)Class I, IINeuroectodermal tumorPhase IValproate and Etoposide for Patients With Neuronal Tumors and Brain Metastases
Brain metastases
Advanced cancer
Valproic AcidiHDAC (fatty acid derivates)Class I, IIBrain tumorsPhase IIValproic Acid With Temozolomide and Radiation Therapy to Treat Brain Tumors
High grade gliomas
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACBrain cancerPhase I/IIPhase I/II Vorinostat, Erlotinib and Temozolomide for Recurrent Glioblastoma Multiforme
GBM multiforme
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACAdult giant cell GBMPhase IIVorinostat and Bortezomib in Treating Patients With Progressive, Recurrent Glioblastoma Multiforme
Adult GBM
Adult gliosarcoma
Recurrent adult brain tumor
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACAdult giant cell GBMPhase IIVorinostat in Treating Patients With Progressive or Recurrent Glioblastoma Multiforme
Adult GBM
Adult gliosarcoma
Recurrent adult brain tumor
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACRecurrent GBM multiformePhase IIPh II SAHA and Bevacizumab for Recurrent Malignant Glioma Patients
Malignant glioma
Adult brain tumor
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACBrain cancerPhase IPhase I Vorinostat Concurrent With Stereotactic Radiosurgery (SRS) in Brain Metastases From Non-Small Cell Lung Cancer
Neoplasm metastasis
Lung cancer
Carcinoma, non-small-cell lung
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACBrain tumorPhase I/IISuberoylanilide Hydroxamic Acid (SAHA), Bevacizumab, Daily Temozolomide for Recurrent Malignant Gliomas
GBM
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACBrain metastasesPhase IStudy of the Combination of Vorinostat and Radiation Therapy for the Treatment of Patients With Brain Metastases
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACAdult GBMNot applicableMagnetic Resonance Spectroscopy Imaging in Predicting Response to Vorinostat and Temozolomide in Patients With Recurrent or Progressive Glioblastoma
Depression
Recurrent adult brain tumor
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACAdult anaplastic astrocytomaPhase IVorinostat and Temozolomide in Treating Patients With Malignant Gliomas
Adult anaplastic oligodendrogliomaAdult giant cell GBM
Adult GBM
Adult gliosarcomaAdult mixed glioma
Recurrent adult brain neoplasm
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACMedulloblastomaPhase IVorinostat Combined With Isotretinoin and Chemotherapy in Treating Younger Patients With Embryonal Tumors of the Central Nervous System
Pineoblastoma
Supratentorial embryonal tumor
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACBrain metastasisPhase IIVorinostat and Concurrent Whole Brain Radiotherapy for Brain Metastasis
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACAdult anaplastic astrocytomaPhase IHigh-Dose Vorinostat and Fractionated Stereotactic Body Radiation Therapy in Treating Patients With Recurrent Glioma
Adult anaplastic oligodendroglioma
Adult giant cell GBM
Adult GBM
Adult gliosarcoma
Adult mixed glioma
Recurrent adult brain neoplasm
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACGBMPhase IPembrolizumab and Vorinostat Combined With Temozolomide for Newly Diagnosed Glioblastoma
Brain tumor
GBM
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACBrain stem gliomaPhase II/IIIVorinostat, Temozolomide, or Bevacizumab in Combination With Radiation Therapy Followed by Bevacizumab and Temozolomide in Young Patients With Newly Diagnosed High-Grade Glioma
Cerebral astrocytoma
Childhood cerebellar anaplastic astrocytoma
Childhood cerebral anaplastic astrocytoma
Childhood spinal cord neoplasm
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACAdult GBMNot applicableMagnetic Resonance Spectroscopy Imaging in Predicting Response to Vorinostat and Temozolomide in Patients With Recurrent or Progressive Glioblastoma
Depression
Recurrent adult brain tumor
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACAdult anaplastic astrocytomaPhase IVorinostat and Temozolomide in Treating Patients With Malignant Gliomas
Adult anaplastic oligodendroglioma
Adult giant cell GBM
Adult GBM
Adult gliosarcoma
Adult mixed glioma
Recurrent adult brain neoplasm
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACMedulloblastomaPhase IVorinostat Combined With Isotretinoin and Chemotherapy in Treating Younger Patients With Embryonal Tumors of the Central Nervous System
Pineoblastoma
Supratentorial embryonal tumor
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACBrain metastasisPhase IIVorinostat and Concurrent Whole Brain Radiotherapy for Brain Metastasis
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACAdult anaplastic astrocytomaPhase IHigh-Dose Vorinostat and Fractionated Stereotactic Body Radiation Therapy in Treating Patients With Recurrent Glioma
Adult anaplastic oligodendroglioma
Adult giant cell GBM
Adult GBM
Adult gliosarcoma
Adult mixed glioma
Recurrent adult brain neoplasm
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACGBMPhase IPembrolizumab and Vorinostat Combined With Temozolomide for Newly Diagnosed GBM
Brain tumor
GBM
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACBrain stem gliomaPhase II/IIIVorinostat, Temozolomide, or Bevacizumab in Combination With Radiation Therapy Followed by Bevacizumab and Temozolomide in Young Patients With Newly Diagnosed High-Grade Glioma
Cerebral astrocytoma
Childhood cerebellar anaplastic astrocytoma
Childhood cerebral anaplastic astrocytoma
Childhood spinal cord neoplasm
Vorinostat (SAHA)iHDAC (hydroxymic acids)Pan-HDACChildhood atypical teratoid/rhabdoid tumorPhase IVorinostat and Temozolomide in Treating Young Patients With Relapsed or Refractory Primary Brain Tumors or Spinal Cord Tumors
Childhood central nervous system choriocarcinoma
Childhood central nervous system embryonal tumor and other

Vorinostat acts by inducing the acetylation of proteins, including histones and transcription factors at both transcriptional and non-transcriptional levels, leading to different cellular effects[68]. Vorinostat, like other pan-HDACi, has variable activity against HDAC isoenzymes, some of which are important in antitumor response.

Vorinostat shows a different toxicity profile compared to classical chemotherapeutic drugs. In fact, common pan-HDACi side effects include fatigue, nausea/vomiting, anemia, anorexia, increased blood urea, hyperglycemia, and thrombocytopenia. Some of these adverse effects can be routinely managed by physicians, but some others need more careful monitoring[69]. However, side effects are dose-dependent and Vorinostat is effective at very low concentrations[70]. There are several reasons behind the epi-drugs’ toxicity. It is partly due to “on-target” effects, which could be explained with the concept of “pleiotropy”. Specifically, a single target gene could be involved in different signaling and controls multiple phenotypic effects. Another reason is “off-target” effects. Epi-drugs are designed to inhibit aberrant epigenetic enzymes, but it is known that they could also affect other classes of substrates belonging to unintended cellular pathways, at both intracellular and extracellular levels[71].

In medulloblastoma, the administration of epigenetic modifiers such as Vorinostat and Valproic acid shows radiosynergistic action on the proliferation and cloning capacity of three human MB cell lines (DAOY, MEB-Med8a, and D283-Med), offering a new opportunity to treat MB patients[72].

DNA methyltransferase inhibitors

The readout of DNMTs is hypermethylation of the DNA sequence and resulting silencing of gene expression. Aberrant DNA methylation has been associated with different stages of cancers. This evidence supports a rationale for using DNMT inhibitors (DNMTi) in cancer treatment (Figure 1).

There are two different classes of DNMTi: (1) Nucleoside analogues, which could be considered analogues to cytosine and act as a natural substrate for DNMT (e.g., 5-azacytidine); and (2) Nonnucleoside compounds, which inhibit DNA methyltransferase activity through mechanisms other than DNA incorporation.

This second class of DNMTi appears to be less toxic and more stable than nucleoside analogues[5,73,74], and Valente and colleagues have focused on the study of nonnucleoside compounds[75].

The archetypal DNA methyltransferase inhibitor is 5-azacytidine (also known as 5-aza), it is a nucleoside inhibitor that is incorporated into the DNA sequence and covalently bonds to and inactivates DNA methyltransferase. As shown in Table 2, 5-azacytidine has been tested in different phase I trials against various brain cancers, particularly in recurrent brain tumors, GBM, and ependymoma.

Table 2 DNA methyltransferase inhibitors in clinical trial for brain cancer treatment.
Compound
Class
Target
Conditions
Phase of clinical trial
Study title
5-Azacytidine DNMTiPan-DNMTBrain tumor recurrentEarly Phase 1Infusion of 5-Azacytidine (5-AZA) Into the Fourth Ventricle in Children With Recurrent Posterior Fossa Ependymoma
AzacitidineDNMTiPan-DNMTGBM multiformeOtherPhase 1Bioequivalence & Food Effect Study in Patients With Solid Tumor or Hematologic Malignancies
5-AzacytidineDNMTiPan-DNMTRecurrent childhood CNS tumorPhase 1Treatment of Children With Recurrent Refractory Brain/Solid Tumors and Recurrent Ependymoma
Ependymoma, recurrent childhood
Childhood solid tumor

Several studies have also demonstrated the differentiation potential of DNMTi in cancer therapy. Liao et al[76] showed that decitabine, an analogue of cytosine, in combinatorial treatment with a differentiation drug, could provide an effective route to enhancing cell differentiation (oligodendrocyte-like morphology and mRNA expression of terminal differentiation marker MBP) and inhibiting cell growth in two malignant glioma human cell lines. Andrade et al[77] tested the efficacy of zebularine, another DNMTi, in four pediatric SHH-MB cell lines (DAOY, ONS-76, UW402, and UW473). Zebularine decreased MB cell growth by targeting the Sonic Hedgehog pathway’s components (GLI1, SMO and PTCH1), evaluated at transcriptional levels. This provides a rationale for further in vivo investigation into the combination of zebularine with chemotherapy[77]. Valente et al[75] tested two nonnucleosides DNMTi (compounds 2 and 5) in mouse MB stem cells isolated from fresh tumor specimens from Ptch1+/- mouse models; compound 2 significantly blocked cell proliferation, while compound 5 was stronger in differentiation potential evaluated by both βIII-tubulin reverse transcriptase polymerase chain reaction and morphology images.

CONCLUSION

Aberrant epigenetic regulation has emerged as a key player in the genesis and progression of brain tumors, influencing malignant phenotypes at various stages of the disease and possibly underlying individual variability in drug response. In particular, the epigenome of brain tumors can be modulated by both cell-intrinsic (e.g., mutations) and cell-extrinsic (e.g., microenvironment) mechanisms, favoring those characteristics of CSCs responsible for cancer growth and disease progression. Reprogramming the epigenetic landscape in the cancer epigenome is among the most promising target therapies, both as a treatment itself and for reversing drug resistance. In this review, we discussed how epigenetic alterations regulate the "stem-like" properties of CSCs and the epigenetic drugs available to blockade epi-mutations. We have reported several examples of epidrugs in Phase I/II clinical trials, providing evidence on the benefit of using epidrugs as single agents or in combination-therapy in brain tumors. We believe this is thus a viable avenue for clinical trials aiming at the development of more affordable and efficient anticancer drugs and treatments.

ACKNOWLEDGEMENTS

We thank Ralf Mouthaan for manuscript editing.

Footnotes

Manuscript source: Invited manuscript

Specialty type: Oncology

Country/Territory of origin: Italy

Peer-review report’s scientific quality classification

Grade A (Excellent): 0

Grade B (Very good): B

Grade C (Good): 0

Grade D (Fair): 0

Grade E (Poor): 0

P-Reviewer: Velasco-Velazquez M S-Editor: Gong ZM L-Editor: Filipodia P-Editor: Xing YX

References
1.  Howlader N, Noone AM, Krapcho M, Miller D, Brest A, Yu M, Ruhl J, Tatalovich Z, Mariotto A, Lewis DR, Chen HS, Feuer EJ, Cronin KA (eds).   SEER Cancer Statistics Review 1975-2017. National Cancer Institute. Bethesda, MD, based on November 2019 SEER data submission, posted to the SEER web site, April 2020. Available from: https://seer.cancer.gov/csr/1975_2017/.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Pollack IF, Agnihotri S, Broniscer A. Childhood brain tumors: current management, biological insights, and future directions. J Neurosurg Pediatr. 2019;23:261-273.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 113]  [Cited by in F6Publishing: 148]  [Article Influence: 29.6]  [Reference Citation Analysis (0)]
3.  Mackay A, Burford A, Carvalho D, Izquierdo E, Fazal-Salom J, Taylor KR, Bjerke L, Clarke M, Vinci M, Nandhabalan M, Temelso S, Popov S, Molinari V, Raman P, Waanders AJ, Han HJ, Gupta S, Marshall L, Zacharoulis S, Vaidya S, Mandeville HC, Bridges LR, Martin AJ, Al-Sarraj S, Chandler C, Ng HK, Li X, Mu K, Trabelsi S, Brahim DH, Kisljakov AN, Konovalov DM, Moore AS, Carcaboso AM, Sunol M, de Torres C, Cruz O, Mora J, Shats LI, Stavale JN, Bidinotto LT, Reis RM, Entz-Werle N, Farrell M, Cryan J, Crimmins D, Caird J, Pears J, Monje M, Debily MA, Castel D, Grill J, Hawkins C, Nikbakht H, Jabado N, Baker SJ, Pfister SM, Jones DTW, Fouladi M, von Bueren AO, Baudis M, Resnick A, Jones C. Integrated Molecular Meta-Analysis of 1,000 Pediatric High-Grade and Diffuse Intrinsic Pontine Glioma. Cancer Cell. 2017;32:520-537.e5.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 669]  [Cited by in F6Publishing: 621]  [Article Influence: 88.7]  [Reference Citation Analysis (0)]
4.  Magee JA, Piskounova E, Morrison SJ. Cancer stem cells: impact, heterogeneity, and uncertainty. Cancer Cell. 2012;21:283-296.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 842]  [Cited by in F6Publishing: 842]  [Article Influence: 70.2]  [Reference Citation Analysis (0)]
5.  Ghasemi S. Cancer's epigenetic drugs: where are they in the cancer medicines? Pharmacogenomics J. 2020;20:367-379.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 39]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
6.  Baumann M, Krause M, Hill R. Clonogens and cancer stem cells. Nat Rev Cancer. 2001;8:990.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Silver DJ, Lathia JD. Revealing the glioma cancer stem cell interactome, one niche at a time. J Pathol. 2018;244:260-264.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 27]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
8.  Bonnet D, Dick JE. Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat Med. 1997;3:730-737.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4851]  [Cited by in F6Publishing: 4671]  [Article Influence: 173.0]  [Reference Citation Analysis (1)]
9.  Ignatova TN, Kukekov VG, Laywell ED, Suslov ON, Vrionis FD, Steindler DA. Human cortical glial tumors contain neural stem-like cells expressing astroglial and neuronal markers in vitro. Glia. 2002;39:193-206.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 729]  [Cited by in F6Publishing: 693]  [Article Influence: 31.5]  [Reference Citation Analysis (0)]
10.  Singh SK, Clarke ID, Terasaki M, Bonn VE, Hawkins C, Squire J, Dirks PB. Identification of a cancer stem cell in human brain tumors. Cancer Res. 2003;63:5821-5828.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Singh SK, Hawkins C, Clarke ID, Squire JA, Bayani J, Hide T, Henkelman RM, Cusimano MD, Dirks PB. Identification of human brain tumour initiating cells. Nature. 2004;432:396-401.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5422]  [Cited by in F6Publishing: 5374]  [Article Influence: 268.7]  [Reference Citation Analysis (0)]
12.  Galli R, Binda E, Orfanelli U, Cipelletti B, Gritti A, De Vitis S, Fiocco R, Foroni C, Dimeco F, Vescovi A. Isolation and characterization of tumorigenic, stem-like neural precursors from human glioblastoma. Cancer Res. 2004;64:7011-7021.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1875]  [Cited by in F6Publishing: 1858]  [Article Influence: 92.9]  [Reference Citation Analysis (0)]
13.  Taylor MD, Poppleton H, Fuller C, Su X, Liu Y, Jensen P, Magdaleno S, Dalton J, Calabrese C, Board J, Macdonald T, Rutka J, Guha A, Gajjar A, Curran T, Gilbertson RJ. Radial glia cells are candidate stem cells of ependymoma. Cancer Cell. 2005;8:323-335.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 652]  [Cited by in F6Publishing: 597]  [Article Influence: 31.4]  [Reference Citation Analysis (0)]
14.  Po A, Ferretti E, Miele E, De Smaele E, Paganelli A, Canettieri G, Coni S, Di Marcotullio L, Biffoni M, Massimi L, Di Rocco C, Screpanti I, Gulino A. Hedgehog controls neural stem cells through p53-independent regulation of Nanog. EMBO J. 2010;29:2646-2658.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 164]  [Cited by in F6Publishing: 174]  [Article Influence: 12.4]  [Reference Citation Analysis (0)]
15.  Manoranjan B, Wang X, Hallett RM, Venugopal C, Mack SC, McFarlane N, Nolte SM, Scheinemann K, Gunnarsson T, Hassell JA, Taylor MD, Lee C, Triscott J, Foster CM, Dunham C, Hawkins C, Dunn SE, Singh SK. FoxG1 interacts with Bmi1 to regulate self-renewal and tumorigenicity of medulloblastoma stem cells. Stem Cells. 2013;31:1266-1277.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 45]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
16.  Tirosh I, Venteicher AS, Hebert C, Escalante LE, Patel AP, Yizhak K, Fisher JM, Rodman C, Mount C, Filbin MG, Neftel C, Desai N, Nyman J, Izar B, Luo CC, Francis JM, Patel AA, Onozato ML, Riggi N, Livak KJ, Gennert D, Satija R, Nahed BV, Curry WT, Martuza RL, Mylvaganam R, Iafrate AJ, Frosch MP, Golub TR, Rivera MN, Getz G, Rozenblatt-Rosen O, Cahill DP, Monje M, Bernstein BE, Louis DN, Regev A, Suvà ML. Single-cell RNA-seq supports a developmental hierarchy in human oligodendroglioma. Nature. 2016;539:309-313.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 641]  [Cited by in F6Publishing: 669]  [Article Influence: 83.6]  [Reference Citation Analysis (0)]
17.  Kim WT, Ryu CJ. Cancer stem cell surface markers on normal stem cells. BMB Rep. 2017;50:285-298.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 156]  [Cited by in F6Publishing: 199]  [Article Influence: 33.2]  [Reference Citation Analysis (0)]
18.  Beier D, Hau P, Proescholdt M, Lohmeier A, Wischhusen J, Oefner PJ, Aigner L, Brawanski A, Bogdahn U, Beier CP. CD133(+) and CD133(-) glioblastoma-derived cancer stem cells show differential growth characteristics and molecular profiles. Cancer Res. 2007;67:4010-4015.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 850]  [Cited by in F6Publishing: 819]  [Article Influence: 48.2]  [Reference Citation Analysis (0)]
19.  Singh SK, Venugopal C.   Brain Tumor Stem Cells Methods and Protocols Methods in Molecular Biology 1869. Humana Press, New York: Springer New York, 2019: 1869.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Glumac PM, LeBeau AM. The role of CD133 in cancer: a concise review. Clin Transl Med. 2018;7:18.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 163]  [Cited by in F6Publishing: 219]  [Article Influence: 36.5]  [Reference Citation Analysis (0)]
21.  Cheng JX, Liu BL, Zhang X. How powerful is CD133 as a cancer stem cell marker in brain tumors? Cancer Treat Rev. 2009;35:403-408.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 76]  [Cited by in F6Publishing: 82]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
22.  Miele E, Po A, Mastronuzzi A, Carai A, Besharat ZM, Pediconi N, Abballe L, Catanzaro G, Sabato C, De Smaele E, Canettieri G, Di Marcotullio L, Vacca A, Mai A, Levrero M, Pfister SM, Kool M, Giangaspero F, Locatelli F, Ferretti E. Downregulation of miR-326 and its host gene β-arrestin1 induces pro-survival activity of E2F1 and promotes medulloblastoma growth. Mol Oncol. 2021;15:523-542.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 4]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
23.  Abbaszadegan MR, Bagheri V, Razavi MS, Momtazi AA, Sahebkar A, Gholamin M. Isolation, identification, and characterization of cancer stem cells: A review. J Cell Physiol. 2017;232:2008-2018.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 115]  [Cited by in F6Publishing: 131]  [Article Influence: 18.7]  [Reference Citation Analysis (0)]
24.  Jonasson E, Ghannoum S, Persson E, Karlsson J, Kroneis T, Larsson E, Landberg G, Ståhlberg A. Identification of Breast Cancer Stem Cell Related Genes Using Functional Cellular Assays Combined With Single-Cell RNA Sequencing in MDA-MB-231 Cells. Front Genet. 2019;10:500.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 23]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
25.  Rodriguez-Meira A, Buck G, Clark SA, Povinelli BJ, Alcolea V, Louka E, McGowan S, Hamblin A, Sousos N, Barkas N, Giustacchini A, Psaila B, Jacobsen SEW, Thongjuea S, Mead AJ. Unravelling Intratumoral Heterogeneity through High-Sensitivity Single-Cell Mutational Analysis and Parallel RNA Sequencing. Mol Cell. 2019;73:1292-1305.e8.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 208]  [Cited by in F6Publishing: 171]  [Article Influence: 34.2]  [Reference Citation Analysis (0)]
26.  Akbarzadeh M, Maroufi NF, Tazehkand AP, Akbarzadeh M, Bastani S, Safdari R, Farzane A, Fattahi A, Nejabati HR, Nouri M, Samadi N. Current approaches in identification and isolation of cancer stem cells. J Cell Physiol. 2019;.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 46]  [Cited by in F6Publishing: 52]  [Article Influence: 10.4]  [Reference Citation Analysis (0)]
27.  Chen J, Li Y, Yu TS, McKay RM, Burns DK, Kernie SG, Parada LF. A restricted cell population propagates glioblastoma growth after chemotherapy. Nature. 2012;488:522-526.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1518]  [Cited by in F6Publishing: 1612]  [Article Influence: 134.3]  [Reference Citation Analysis (0)]
28.  Bao S, Wu Q, McLendon RE, Hao Y, Shi Q, Hjelmeland AB, Dewhirst MW, Bigner DD, Rich JN. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature. 2006;444:756-760.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4467]  [Cited by in F6Publishing: 4528]  [Article Influence: 251.6]  [Reference Citation Analysis (0)]
29.  Elsässer SJ, Allis CD, Lewis PW. Cancer. New epigenetic drivers of cancers. Science. 2011;331:1145-1146.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 62]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
30.  Northcott PA, Lee C, Zichner T, Stütz AM, Erkek S, Kawauchi D, Shih DJ, Hovestadt V, Zapatka M, Sturm D, Jones DT, Kool M, Remke M, Cavalli FM, Zuyderduyn S, Bader GD, VandenBerg S, Esparza LA, Ryzhova M, Wang W, Wittmann A, Stark S, Sieber L, Seker-Cin H, Linke L, Kratochwil F, Jäger N, Buchhalter I, Imbusch CD, Zipprich G, Raeder B, Schmidt S, Diessl N, Wolf S, Wiemann S, Brors B, Lawerenz C, Eils J, Warnatz HJ, Risch T, Yaspo ML, Weber UD, Bartholomae CC, von Kalle C, Turányi E, Hauser P, Sanden E, Darabi A, Siesjö P, Sterba J, Zitterbart K, Sumerauer D, van Sluis P, Versteeg R, Volckmann R, Koster J, Schuhmann MU, Ebinger M, Grimes HL, Robinson GW, Gajjar A, Mynarek M, von Hoff K, Rutkowski S, Pietsch T, Scheurlen W, Felsberg J, Reifenberger G, Kulozik AE, von Deimling A, Witt O, Eils R, Gilbertson RJ, Korshunov A, Taylor MD, Lichter P, Korbel JO, Wechsler-Reya RJ, Pfister SM. Enhancer hijacking activates GFI1 family oncogenes in medulloblastoma. Nature. 2014;511:428-434.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 414]  [Cited by in F6Publishing: 431]  [Article Influence: 43.1]  [Reference Citation Analysis (0)]
31.  Feinberg AP, Tycko B. The history of cancer epigenetics. Nat Rev Cancer. 2004;4:143-153.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1674]  [Cited by in F6Publishing: 1498]  [Article Influence: 74.9]  [Reference Citation Analysis (0)]
32.  Alelú-Paz R, Ashour N, González-Corpas A, Ropero S. DNA methylation, histone modifications, and signal transduction pathways: a close relationship in malignant gliomas pathophysiology. J Signal Transduct. 2012;2012:956958.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 16]  [Cited by in F6Publishing: 16]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
33.  Lujambio A, Portela A, Liz J, Melo SA, Rossi S, Spizzo R, Croce CM, Calin GA, Esteller M. CpG island hypermethylation-associated silencing of non-coding RNAs transcribed from ultraconserved regions in human cancer. Oncogene. 2010;29:6390-6401.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 147]  [Cited by in F6Publishing: 156]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
34.  Ellis L, Atadja PW, Johnstone RW. Epigenetics in cancer: targeting chromatin modifications. Mol Cancer Ther. 2009;8:1409-1420.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 339]  [Cited by in F6Publishing: 323]  [Article Influence: 21.5]  [Reference Citation Analysis (0)]
35.  Takeshima H, Ushijima T.   DNA methylation changes in cancer: Mechanisms. In: Boffetta P, Hainaut P. Encyclopedia of Cancer. 3rd ed. Elsevier, 2018: 520-529.  [PubMed]  [DOI]  [Cited in This Article: ]
36.  Ohm JE, McGarvey KM, Yu X, Cheng L, Schuebel KE, Cope L, Mohammad HP, Chen W, Daniel VC, Yu W, Berman DM, Jenuwein T, Pruitt K, Sharkis SJ, Watkins DN, Herman JG, Baylin SB. A stem cell-like chromatin pattern may predispose tumor suppressor genes to DNA hypermethylation and heritable silencing. Nat Genet. 2007;39:237-242.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 837]  [Cited by in F6Publishing: 809]  [Article Influence: 47.6]  [Reference Citation Analysis (0)]
37.  Feinberg AP, Vogelstein B. Hypomethylation distinguishes genes of some human cancers from their normal counterparts. Nature. 1983;301:89-92.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1699]  [Cited by in F6Publishing: 1565]  [Article Influence: 38.2]  [Reference Citation Analysis (0)]
38.  Toraño EG, Petrus S, Fernandez AF, Fraga MF. Global DNA hypomethylation in cancer: review of validated methods and clinical significance. Clin Chem Lab Med. 2012;50:1733-1742.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 70]  [Article Influence: 5.8]  [Reference Citation Analysis (0)]
39.  Gopisetty G, Xu J, Sampath D, Colman H, Puduvalli VK. Epigenetic regulation of CD133/PROM1 expression in glioma stem cells by Sp1/myc and promoter methylation. Oncogene. 2013;32:3119-3129.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 53]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
40.  Easwaran H, Tsai HC, Baylin SB. Cancer epigenetics: tumor heterogeneity, plasticity of stem-like states, and drug resistance. Mol Cell. 2014;54:716-727.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 607]  [Cited by in F6Publishing: 649]  [Article Influence: 64.9]  [Reference Citation Analysis (0)]
41.  Capper D, Jones DTW, Sill M, Hovestadt V, Schrimpf D, Sturm D, Koelsche C, Sahm F, Chavez L, Reuss DE, Kratz A, Wefers AK, Huang K, Pajtler KW, Schweizer L, Stichel D, Olar A, Engel NW, Lindenberg K, Harter PN, Braczynski AK, Plate KH, Dohmen H, Garvalov BK, Coras R, Hölsken A, Hewer E, Bewerunge-Hudler M, Schick M, Fischer R, Beschorner R, Schittenhelm J, Staszewski O, Wani K, Varlet P, Pages M, Temming P, Lohmann D, Selt F, Witt H, Milde T, Witt O, Aronica E, Giangaspero F, Rushing E, Scheurlen W, Geisenberger C, Rodriguez FJ, Becker A, Preusser M, Haberler C, Bjerkvig R, Cryan J, Farrell M, Deckert M, Hench J, Frank S, Serrano J, Kannan K, Tsirigos A, Brück W, Hofer S, Brehmer S, Seiz-Rosenhagen M, Hänggi D, Hans V, Rozsnoki S, Hansford JR, Kohlhof P, Kristensen BW, Lechner M, Lopes B, Mawrin C, Ketter R, Kulozik A, Khatib Z, Heppner F, Koch A, Jouvet A, Keohane C, Mühleisen H, Mueller W, Pohl U, Prinz M, Benner A, Zapatka M, Gottardo NG, Driever PH, Kramm CM, Müller HL, Rutkowski S, von Hoff K, Frühwald MC, Gnekow A, Fleischhack G, Tippelt S, Calaminus G, Monoranu CM, Perry A, Jones C, Jacques TS, Radlwimmer B, Gessi M, Pietsch T, Schramm J, Schackert G, Westphal M, Reifenberger G, Wesseling P, Weller M, Collins VP, Blümcke I, Bendszus M, Debus J, Huang A, Jabado N, Northcott PA, Paulus W, Gajjar A, Robinson GW, Taylor MD, Jaunmuktane Z, Ryzhova M, Platten M, Unterberg A, Wick W, Karajannis MA, Mittelbronn M, Acker T, Hartmann C, Aldape K, Schüller U, Buslei R, Lichter P, Kool M, Herold-Mende C, Ellison DW, Hasselblatt M, Snuderl M, Brandner S, Korshunov A, von Deimling A, Pfister SM. DNA methylation-based classification of central nervous system tumours. Nature. 2018;555:469-474.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1746]  [Cited by in F6Publishing: 1568]  [Article Influence: 261.3]  [Reference Citation Analysis (0)]
42.  Hosseini A, Minucci S.   Alterations of Histone Modifications in Cancer. In: Tollefsbol TO. In Translational Epigenetics, Epigenetics in Human Disease. 2nd ed. Elsevier Inc, 2018: 141-217.  [PubMed]  [DOI]  [Cited in This Article: ]
43.  Zhou D, Alver BM, Li S, Hlady RA, Thompson JJ, Schroeder MA, Lee JH, Qiu J, Schwartz PH, Sarkaria JN, Robertson KD. Distinctive epigenomes characterize glioma stem cells and their response to differentiation cues. Genome Biol. 2018;19:43.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 55]  [Cited by in F6Publishing: 57]  [Article Influence: 9.5]  [Reference Citation Analysis (0)]
44.  McCann TS, Sobral LM, Self C, Hsieh J, Sechler M, Jedlicka P. Biology and targeting of the Jumonji-domain histone demethylase family in childhood neoplasia: a preclinical overview. Expert Opin Ther Targets. 2019;23:267-280.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 10]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
45.  Liau BB, Sievers C, Donohue LK, Gillespie SM, Flavahan WA, Miller TE, Venteicher AS, Hebert CH, Carey CD, Rodig SJ, Shareef SJ, Najm FJ, van Galen P, Wakimoto H, Cahill DP, Rich JN, Aster JC, Suvà ML, Patel AP, Bernstein BE. Adaptive Chromatin Remodeling Drives Glioblastoma Stem Cell Plasticity and Drug Tolerance. Cell Stem Cell. 2017;20:233-246.e7.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 341]  [Cited by in F6Publishing: 322]  [Article Influence: 46.0]  [Reference Citation Analysis (0)]
46.  Marampon F, Megiorni F, Camero S, Crescioli C, McDowell HP, Sferra R, Vetuschi A, Pompili S, Ventura L, De Felice F, Tombolini V, Dominici C, Maggio R, Festuccia C, Gravina GL. HDAC4 and HDAC6 sustain DNA double strand break repair and stem-like phenotype by promoting radioresistance in glioblastoma cells. Cancer Lett. 2017;397:1-11.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 65]  [Article Influence: 9.3]  [Reference Citation Analysis (0)]
47.  Banelli B, Carra E, Barbieri F, Würth R, Parodi F, Pattarozzi A, Carosio R, Forlani A, Allemanni G, Marubbi D, Florio T, Daga A, Romani M. The histone demethylase KDM5A is a key factor for the resistance to temozolomide in glioblastoma. Cell Cycle. 2015;14:3418-3429.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 83]  [Article Influence: 10.4]  [Reference Citation Analysis (0)]
48.  Miele E, Valente S, Alfano V, Silvano M, Mellini P, Borovika D, Marrocco B, Po A, Besharat ZM, Catanzaro G, Battaglia G, Abballe L, Zwergel C, Stazi G, Milite C, Castellano S, Tafani M, Trapencieris P, Mai A, Ferretti E. The histone methyltransferase EZH2 as a druggable target in SHH medulloblastoma cancer stem cells. Oncotarget. 2017;8:68557-68570.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 31]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
49.  Yao Q, Chen Y, Zhou X. The roles of microRNAs in epigenetic regulation. Curr Opin Chem Biol. 2019;51:11-17.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 179]  [Cited by in F6Publishing: 247]  [Article Influence: 49.4]  [Reference Citation Analysis (0)]
50.  Ferretti E, De Smaele E, Miele E, Laneve P, Po A, Pelloni M, Paganelli A, Di Marcotullio L, Caffarelli E, Screpanti I, Bozzoni I, Gulino A. Concerted microRNA control of Hedgehog signalling in cerebellar neuronal progenitor and tumour cells. EMBO J. 2008;27:2616-2627.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 239]  [Cited by in F6Publishing: 261]  [Article Influence: 16.3]  [Reference Citation Analysis (0)]
51.  Lizarte Neto FS, Rodrigues AR, Trevisan FA, de Assis Cirino ML, Matias CCMS, Pereira-da-Silva G, Peria FM, Tirapelli DPDC, Carlotti CG Jr. microRNA-181d associated with the methylation status of the MGMT gene in Glioblastoma multiforme cancer stem cells submitted to treatments with ionizing radiation and temozolomide. Brain Res. 2019;1720:146302.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7]  [Cited by in F6Publishing: 8]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
52.  Yang E, Wang X, Gong Z, Yu M, Wu H, Zhang D. Exosome-mediated metabolic reprogramming: the emerging role in tumor microenvironment remodeling and its influence on cancer progression. Signal Transduct Target Ther. 2020;5:242.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 168]  [Article Influence: 42.0]  [Reference Citation Analysis (0)]
53.  Bao S, Tang F, Li X, Hayashi K, Gillich A, Lao K, Surani MA. Epigenetic reversion of post-implantation epiblast to pluripotent embryonic stem cells. Nature. 2009;461:1292-1295.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 303]  [Cited by in F6Publishing: 312]  [Article Influence: 20.8]  [Reference Citation Analysis (0)]
54.  Downing TL, Soto J, Morez C, Houssin T, Fritz A, Yuan F, Chu J, Patel S, Schaffer DV, Li S. Biophysical regulation of epigenetic state and cell reprogramming. Nat Mater. 2013;12:1154-1162.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 356]  [Cited by in F6Publishing: 344]  [Article Influence: 31.3]  [Reference Citation Analysis (0)]
55.  Citron F, Fabris L. Targeting Epigenetic Dependencies in Solid Tumors: Evolutionary Landscape Beyond Germ Layers Origin. Cancers (Basel). 2020;12:682.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5]  [Cited by in F6Publishing: 5]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
56.  Eckschlager T, Plch J, Stiborova M, Hrabeta J. Histone Deacetylase Inhibitors as Anticancer Drugs. Int J Mol Sci. 2017;18.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 643]  [Cited by in F6Publishing: 778]  [Article Influence: 111.1]  [Reference Citation Analysis (0)]
57.  Perla A, Fratini L, Cardoso PS, Nör C, Brunetto AT, Brunetto AL, de Farias CB, Jaeger M, Roesler R. Histone Deacetylase Inhibitors in Pediatric Brain Cancers: Biological Activities and Therapeutic Potential. Front Cell Dev Biol. 2020;8:546.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 26]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
58.  Reddy RG, Bhat UA, Chakravarty S, Kumar A. Advances in histone deacetylase inhibitors in targeting glioblastoma stem cells. Cancer Chemother Pharmacol. 2020;86:165-179.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 6]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
59.  Staberg M, Michaelsen SR, Rasmussen RD, Villingshøj M, Poulsen HS, Hamerlik P. Inhibition of histone deacetylases sensitizes glioblastoma cells to lomustine. Cell Oncol (Dordr). 2017;40:21-32.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 37]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
60.  Johnstone RW. Histone-deacetylase inhibitors: novel drugs for the treatment of cancer. Nat Rev Drug Discov. 2002;1:287-299.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1146]  [Cited by in F6Publishing: 1116]  [Article Influence: 50.7]  [Reference Citation Analysis (0)]
61.  da Cunha Jaeger M, Ghisleni EC, Cardoso PS, Siniglaglia M, Falcon T, Brunetto AT, Brunetto AL, de Farias CB, Taylor MD, Nör C, Ramaswamy V, Roesler R. HDAC and MAPK/ERK Inhibitors Cooperate To Reduce Viability and Stemness in Medulloblastoma. J Mol Neurosci. 2020;70:981-992.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 19]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
62.  Coni S, Mancuso AB, Di Magno L, Sdruscia G, Manni S, Serrao SM, Rotili D, Spiombi E, Bufalieri F, Petroni M, Kusio-Kobialka M, De Smaele E, Ferretti E, Capalbo C, Mai A, Niewiadomski P, Screpanti I, Di Marcotullio L, Canettieri G. Selective targeting of HDAC1/2 elicits anticancer effects through Gli1 acetylation in preclinical models of SHH Medulloblastoma. Sci Rep. 2017;7:44079.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 46]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
63.  Meel MH, de Gooijer MC, Metselaar DS, Sewing ACP, Zwaan K, Waranecki P, Breur M, Buil LCM, Lagerweij T, Wedekind LE, Twisk JWR, Koster J, Hashizume R, Raabe EH, Montero Carcaboso Á, Bugiani M, Phoenix TN, van Tellingen O, van Vuurden DG, Kaspers GJL, Hulleman E. Combined Therapy of AXL and HDAC Inhibition Reverses Mesenchymal Transition in Diffuse Intrinsic Pontine Glioma. Clin Cancer Res. 2020;26:3319-3332.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 35]  [Article Influence: 8.8]  [Reference Citation Analysis (0)]
64.  Shofuda T, Kanemura Y. HDACs and MYC in medulloblastoma: how do HDAC inhibitors control MYC-amplified tumors? Neuro Oncol. 2021;23:173-174.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 3]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
65.  Orzan F, Pellegatta S, Poliani PL, Pisati F, Caldera V, Menghi F, Kapetis D, Marras C, Schiffer D, Finocchiaro G. Enhancer of Zeste 2 (EZH2) is up-regulated in malignant gliomas and in glioma stem-like cells. Neuropathol Appl Neurobiol. 2011;37:381-394.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 95]  [Cited by in F6Publishing: 104]  [Article Influence: 8.0]  [Reference Citation Analysis (0)]
66.  Hsu CC, Chang WC, Hsu TI, Liu JJ, Yeh SH, Wang JY, Liou JP, Ko CY, Chang KY, Chuang JY. Suberoylanilide hydroxamic acid represses glioma stem-like cells. J Biomed Sci. 2016;23:81.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 28]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
67.  Sung GJ, Kim SH, Kwak S, Park SH, Song JH, Jung JH, Kim H, Choi KC. Inhibition of TFEB oligomerization by co-treatment of melatonin with vorinostat promotes the therapeutic sensitivity in glioblastoma and glioma stem cells. J Pineal Res. 2019;66:e12556.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 24]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
68.  Bubna AK. Vorinostat-An Overview. Indian J Dermatol. 2015;60:419.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 73]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
69.  Subramanian S, Bates SE, Wright JJ, Espinoza-Delgado I, Piekarz RL. Clinical Toxicities of Histone Deacetylase Inhibitors. Pharmaceuticals (Basel). 2010;3:2751-2767.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 214]  [Cited by in F6Publishing: 237]  [Article Influence: 16.9]  [Reference Citation Analysis (0)]
70.  Blumenschein GR Jr, Kies MS, Papadimitrakopoulou VA, Lu C, Kumar AJ, Ricker JL, Chiao JH, Chen C, Frankel SR. Phase II trial of the histone deacetylase inhibitor vorinostat (Zolinza, suberoylanilide hydroxamic acid, SAHA) in patients with recurrent and/or metastatic head and neck cancer. Invest New Drugs. 2008;26:81-87.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 129]  [Cited by in F6Publishing: 113]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
71.  Singh BN, Zhang G, Hwa YL, Li J, Dowdy SC, Jiang SW. Nonhistone protein acetylation as cancer therapy targets. Expert Rev Anticancer Ther. 2010;10:935-954.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 207]  [Cited by in F6Publishing: 213]  [Article Influence: 15.2]  [Reference Citation Analysis (0)]
72.  Patties I, Kortmann RD, Menzel F, Glasow A. Enhanced inhibition of clonogenic survival of human medulloblastoma cells by multimodal treatment with ionizing irradiation, epigenetic modifiers, and differentiation-inducing drugs. J Exp Clin Cancer Res. 2016;35:94.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14]  [Cited by in F6Publishing: 15]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
73.  Lyko F, Brown R. DNA methyltransferase inhibitors and the development of epigenetic cancer therapies. J Natl Cancer Inst. 2005;97:1498-1506.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 357]  [Cited by in F6Publishing: 347]  [Article Influence: 18.3]  [Reference Citation Analysis (0)]
74.  Zwergel C, Valente S, Mai A. DNA Methyltransferases Inhibitors from Natural Sources. Curr Top Med Chem. 2016;16:680-696.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 44]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
75.  Valente S, Liu Y, Schnekenburger M, Zwergel C, Cosconati S, Gros C, Tardugno M, Labella D, Florean C, Minden S, Hashimoto H, Chang Y, Zhang X, Kirsch G, Novellino E, Arimondo PB, Miele E, Ferretti E, Gulino A, Diederich M, Cheng X, Mai A. Selective non-nucleoside inhibitors of human DNA methyltransferases active in cancer including in cancer stem cells. J Med Chem. 2014;57:701-713.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 80]  [Cited by in F6Publishing: 89]  [Article Influence: 8.9]  [Reference Citation Analysis (0)]
76.  Zaba R. [Breath test using C-13-trioleate in the evaluation of the rate of fatty acid metabolism after parenteral feeding of premature and newborn infants]. Pediatr Pol. 1988;63:661-664.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 11]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
77.  Andrade AF, Borges KS, Suazo VK, Geron L, Corrêa CA, Castro-Gamero AM, de Vasconcelos EJ, de Oliveira RS, Neder L, Yunes JA, Dos Santos Aguiar S, Scrideli CA, Tone LG. The DNA methyltransferase inhibitor zebularine exerts antitumor effects and reveals BATF2 as a poor prognostic marker for childhood medulloblastoma. Invest New Drugs. 2017;35:26-36.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 16]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]