We use cookies to improve your experience. By continuing to browse this site, you accept our cookie policy.×
Skip main navigation
Aging Health
Bioelectronics in Medicine
Biomarkers in Medicine
Breast Cancer Management
CNS Oncology
Colorectal Cancer
Concussion
Epigenomics
Future Cardiology
Future Medicine AI
Future Microbiology
Future Neurology
Future Oncology
Future Rare Diseases
Future Virology
Hepatic Oncology
HIV Therapy
Immunotherapy
International Journal of Endocrine Oncology
International Journal of Hematologic Oncology
Journal of 3D Printing in Medicine
Lung Cancer Management
Melanoma Management
Nanomedicine
Neurodegenerative Disease Management
Pain Management
Pediatric Health
Personalized Medicine
Pharmacogenomics
Regenerative Medicine
ReviewOpen Accesscc iconby icon

CRISPR-based epigenome editing: mechanisms and applications

    Shaima M Fadul

    Department of Life Sciences, College of Science & General Studies, Alfaisal University, Riyadh, 11533, Kingdom of Saudi Arabia

    ,
    Aleeza Arshad

    Medical Teaching Insitute, Ayub Teaching Hospital, Abbottabad, 22020, Pakistan

    &
    Rashid Mehmood

    *Author for correspondence:

    E-mail Address: rmehmood@alfaisal.edu

    Department of Life Sciences, College of Science & General Studies, Alfaisal University, Riyadh, 11533, Kingdom of Saudi Arabia

    Published Online:https://doi.org/10.2217/epi-2023-0281

    Abstract

    Epigenomic anomalies contribute significantly to the development of numerous human disorders. The development of epigenetic research tools is essential for understanding how epigenetic marks contribute to gene expression. A gene-editing technique known as CRISPR (clustered regularly interspaced short palindromic repeats) typically targets a particular DNA sequence using a guide RNA (gRNA). CRISPR/Cas9 technology has been remodeled for epigenome editing by generating a ‘dead’ Cas9 protein (dCas9) that lacks nuclease activity and juxtaposing it with an epigenetic effector domain. Based on fusion partners of dCas9, a specific epigenetic state can be achieved. CRISPR-based epigenome editing has widespread application in drug screening, cancer treatment and regenerative medicine. This paper discusses the tools developed for CRISPR-based epigenome editing and their applications.

    Tweetable abstract

    CRISPR/Cas9 system has been remodeled for epigenome editing by juxtaposing ‘dead’ dCas9 with an epigenetic effector domain. This tool has applications in cancer treatment and regenerative medicine.

    “The complex developmental processes that connect between the genotype and the phenotype” is typically recognized as the first definition of epigenetics established by Conrad Waddington [1]. Although the term has been refined since then, it remains an accurate description of epigenetics. It encompasses a plethora of mechanisms that influence the phenotype beyond just the DNA sequence. Such mechanisms that regulate genomic functions are mediated by epigenetic factors. These factors effectuate various modifications that include DNA methylation, histone modifications [2–4], chromatin accessibility [5], 3D genome organization [6], and interactions of various RNA and protein elements with the genome (Figure 1) [7–10].

    Figure 1. Molecular mechanisms modulating the epigenome.

    The epigenome encompasses a variety of dynamic molecular processes crucial for cell function and development. (A) Nucleosomes, composed of DNA (black line) wrapped around histones (blue) are the basic units of chromatin packaging in eukaryotes. Histones tails can be modified by the covalent additions of chemical groups catalyzed by enzymes (colored circles), ultimately effecting chromatin accessibility. (B) Methyltransferases may add a methyl group to the C-5 position of cytosine, which may affect transcription status. (C) Noncoding RNA elements such as lncRNAs (red line) can modulate gene expression by binding TFs (yellow circle). (D) Chromatin is intricately folded into a 3D configuration in the nucleus, constituting chromatin domains and interactions of regulatory elements that affect gene expression.

    lncRNA: Long noncoding RNA; TF: Transcription factor.

    These factors orchestrate to modify the epigenetic landscape of the cell and determine which genes are turned ‘on’ and ‘off’ to control gene expression and ultimately cell fate. Genome regulation by the epigenomic mechanisms ensures that each cell is only able to produce molecules necessary for its cell or tissue function [11–13]. Epigenetic factors not only play a role in cell fate determination but also in the propagation and maintenance of the epigenetic state of the cells. Maintaining cellular memory across cell divisions is essential to maintain normal cell and organ function [14,15].

    Owing to the significant role of epigenetics in normal physiology, it is not surprising that the epigenetic landscape is altered in a number of human disorders [16] including cancer [17–19], heart disease [20–22], obesity [23,24], neurological disorders [25,26], and so on. Understanding how epigenetic factors individually and collectively contribute to modulation of the epigenetic landscape, and how aberrant epigenetic profiles promote disease progression, requires the development and expansion of epigenetics research tools. The utilization of such tools will be valuable in both basic and applied research [16]. Identifying aberrant epigenetic patterns may serve as biomarkers in diagnostics [27,28], and in the development of epigenetic drugs called ‘epidrugs’ that target epigenetic mechanisms in cancer and other human disorders [27,29]. Despite their potential use in a number of conditions, the nonspecificity of epidrugs poses a serious challenge. Recently, several biochemical and bioinformatic tools have been developed to study and alter epigenetic marks, mostly focusing on methylation and histone modifications. However, these techniques have some limitations in terms of their versatility and precision [30,31]. Therefore, it is critical to develop novel technologies that are precise in inducing and interfering with epigenetic marks at the specific genomic regions. One such tools is a modified version of CRISPR-Cas system.

    In this paper we will discuss the tools developed for CRISPR-based epigenome editing and their applications. We will begin by discussing the basic principles of CRISPR and how it can be utilized to target the epigenetic landscape. We then discuss the different tools that have been developed for epigenome modulation, and lastly review the current state of knowledge on the applications of CRISPR systems in this field.

    Harnessing CRISPR for epigenome editing

    the CRISPR system, originally identified in bacteria and archea, is a revolutionary gene-editing tool that typically uses guide RNA (gRNA) to target a specific DNA sequence. Once the gRNA binds to the target DNA sequence, Cas proteins are recruited to the site and cut the DNA to allow insertions, deletions and modifications on the target site using indigenous cell DNA-repair mechanisms [32]. The CRISPR-Cas system exhibit high functional and structural diversity. CRISPR-Cas can be categorized into six different types (I–VI), each type consisting of a distinct Cas nuclease protein. The six different types have been further regrouped into two classes depending on the composition of Cas effectors and CRISPR RNA ribonucleoprotein (crRNP) complexes.

    Class 1 systems include (Types III, IV, and I) and consist of multiple Cas proteins in their crRNPs that can include nuclease components. While Class 2 systems (types II, V, and VI) only consist of a single Cas effector Protein in their crRNP complexes [33,34]. This diversity also includes the nucleic acid substrates that can be recognized and targeted by CRISPR-Cas effector Complexes. Types I, II, V and likely IV target DNA, while type III systems can target both DNA and RNA [35]. Type VI systems exclusively target RNA [36,37]. CRISPR system research tools were originally developed utilizing class 2 type II Cas9 and type V Cas12 single effector complexes that originally target and cleave DNA sequences. The CRISPR-based toolbox expanded to include nuclease ‘null’ Cas proteins dCas9 and dCas12 [38]. A concerted effort is now underway to include CRISPR systems that bind and cleave RNA molecules based on type III and VI crRNP effector complexes [35,39–42].

    Recently, CRISPR/Cas9 technology has been remolded for epigenome editing by generating a ‘dead’ Cas9 protein (dCas9) or ‘nuclease-null’ that allows us to take advantage of its DNA-binding abilities to target specific loci without cleavage [43]. Epigenetic effectors can be tethered to the catalytically inactive Cas9 complex, allowing it to localize the effectors to a specific locus to induce epigenetic alterations. The same dCas9 protein can be reused to target other loci just by changing the sgRNA sequence. CRISPR technology offers great advantages as an engineering tool from targeting DNA sequences of any length, to specific and multiplex modifications [44–46]. The developments of dCas9 fusion constructs provided us with a diverse epigenome control toolkit, to control transcriptional activation and repression of genes at the level of histone modifications [47,48], DNA methylation [49,50], and even of genes coding for noncoding RNAs (ncRNAs) [51,52]. Table 1 summarizes the tools developed from various classes of CRISPR systems. In the following section, we will discuss the mechanisms that have been employed to design tools for CRISPR-based epigenome editing and their applications (summarized in Table 2).

    Table 1. Types of CRISPR systems that have been repurposed for epigenome editing.
    TypeIIIIIIIVVVIRef.
    Class1 (multisubunit)2 (single unit)1 (multisubunit)1 (multisubunit)2 (single unit)2 (single unit)
    Endonuclease proteinCas3Cas9Cas10Scf1Cas12Cas13 
    Target moleculeDNADNADNA/RNADNARNA 
    Epigenome editing platformsdCas9-DNMT3A
    dCas9-SunTag-DNMT3A
    dCas9-DNMT3A-DNMT3L
    dCas9-DNMT3A-DNMT3L-KRAB
    dCas9-TET1CD
    dCas9-Tet1-CD and MS2-Tet1-CD dual system
    dCas9-TET1
    TET1-dSaCas9 and VPR-dSpCas9
    dCas9-SunTag-TET1
    dCas9-p300
    dCas9-HDAC3
    dCas9-KRAB
    dCas9-LSD1
    dCas9-Ezh2
    dCas9-VPR (CRISPRa)
    DiCas7-11CRISPR-Cas13d
    dCas13b-ADAR2
    CRISPR-Cas13a
    [36,46,47,49,53–76]
    Table 2. dCas9-mediated epigenome editing involves multiple epigenetic mechanisms.
    ModificationEffector domainsTargetMechanismRef.
    DNA methylationdCas9-DNMT3A
    dCas9-SunTag-DNMT3A
    dCas9-DNMT3A-DNMT3L
    dCas9-DNMT3A-DNMT3L-KRAB
    BDNF promoter
    MyoD enhancer
    CTCF loops
    CDKN2A promoter, ARF promoter, Cdkn1a promoter
    IL6ST, BACH2
    HOXA5
    EpCAM, CXCR4 and TFRC promoters
    UNC5C promoter
    GAPDH-Snrpn reporter system
    Targeting of dCas9-DNMT3a fusion proteins to unmethylated promoter sequences
    Targeting of the dCas9-DNMT3a fusion domain along with repetitive peptide epitopes (SunTag) that can recruit and amplify the local concentration of DNMT3a domains for robust methylation of target loci
    Targeting of a dCas9-DNMT3A-DNMT3L methyltransferase chimeric construct to induce higher methylation efficiency compared with a system that utilizes a single methyltransferase catalytic domain
    Development of ‘CRISPRoff’ composed of dCas9-DNMT3A-DNMT3L-KRAB fusion domains to target and induce gene silencing memory
    [46,49,56–59,62,77]
    DNA demethylationdCas9-TET1CD
    dCas9-Tet1-CD and MS2-Tet1-CD dual system
    dCas9-TET1
    TET1-dSaCas9 and VPR-dSpCas9
    dCas9-SunTag-TET1
    enzyme free steric interference with DNA methyltransferases
    dCas9-TETv4 with sgRNA modified with MCP (VP64, p65-AD and Rta)
    STAT3 binding site upstream of GFAP
    BRCA1 gene promoter
    RANKL, MAGEB2, MMP2 neighboring CpGs
    BDNF promoter
    MyoD enhancer
    Dazl-Snrpn-GFP reporter
    HNF1A, MGAT3
    FMR1, SerpinB5, Tnf
    CLTA
    Targeted DNA demethylation using a dCas9 fusion construct with a TET1CD catalytic domain
    Simultaneous transfection of dCas9-TET1-CD and a MS2 coat protein fusion construct guided with a sgRNA2.0 transcription system to demethylate target genomic regions and promote gene expression
    Induction of de novo demethylation of target gene regions with the dCas9-TET1 construct
    Demethylation of regulatory regions via dual TET1-dSaCas9 and VPR-dSpCas9 fusion constructs for synergistic modulation of gene expression
    Targeting of the dCas9-TET1 fusion domain along with repetitive peptide epitopes (SunTag) that can recruit and amplify the local concentration of TET1 proteins for robust demethylation of target loci
    Targeting of the enzyme-free dCas9 system to CpGs adjacent to transcription start sites to interfere with the binding of DNA methyltransferases leading to induced site-specific gene expression
    Development of ‘CRISPRon’ composed of a dCas9-TETv4 fusion domain with modified sgRNAs with MCP and various combinations of transactivation domains to reverse ‘CRISPRoff’-mediated gene silencing
    [46,53–57,60,61,77]
    H3K27 acetylationdCas9-p300
    dCas9-p300 fused with PYL or ABI proteins
    dLbCpf1-P300/dCpf1-SunTag
    dCas9-p300 and MCP-VP64 dual system
    IL1RN (MYOD), POU5F1/OCT4
    IL1RN
    GRM2
    MYOD, HS2 enhancer, TAL1
    IL1RN, GRM2, HBA, MYOD
    IL1RN, RHOXF2, TTN
    Targeting of dCas9-p300 acetyltransferase domain to promoters and distal enhancers leading to gene activation
    Utilizing CIP for temporal control of histone modifications, where p300 and dCas9 proteins can be fused to ABI and PYL that can dimerize when the small chemical inducer ABA is introduced to reversibly recruit p300 to target sites
    System based on a Cpf1 and p300 fusion domain for targeted histone acetylation as an alternative to dCas9 to activate multiple genes with a single sgRNA
    enhancer targeting by ‘enCRISPRa’ system for activation
    [47,63–65,78–82]
    H3K27 deacetylationdCas9-HDAC3
    dCas9-KRAB
    Smn1, Mecp2, Isl1
    OCT4, Tbx3, SOX2
    Targeting of the dCas9-HDAC3 deacetylase domain to promoters and enhancers leading to a repressive state
    Targeting of the dCas9-KRAB deacetylase domain to promoters and enhancers leading to a repressive state
    [66,67]
    H3K4 demethylationdCas9-LSD1OCT4, Tbx3, SOX2Targeting of the dCas9-LSD1 demethylase domain to promoters and enhancers leading to activation[67]
    H3K4 methylationdCas9-SETICAM1, RASSF1a, EpCAM
    IL1RN, GRM2
    Targeting of the dCas9-SET histone methyltransferase domain to promoters leading to gene reactivation[63,83]
    H3K9 methylationdCas9-G9A[SET]/dCas9-SUV[SET]
    dCas9-KRAB
    HER2, MYC, EPCAM
    HS2 enhancer
    Repression of target genes by histone methyltransferases fusion domains with dCas9
    Targeting of dCas9-KRAB to enhancer regions and recruitment of repressive chromatin methylation modifiers
    [68,84]
    H3K27 methylationdCas9-olEzh2
    dCas9-Ezh2
    Arhgap35, Pfkfb4a, Nanos3, Dcx, Kita, Tbx16, Slc41a2a
    HER2, MYC, EPCAM

    C/ebpa promoter
    Repression of target genes by histone methyltransferases fusion domains with dCas9[68,69,85]
    RNADECKO (double-excision CRISPR knockout)
    CRISPR-Cas13d
    dCas9-VPR (CRISPRa)
    dCas13b-ADAR2
    DiCas7-11
    CRISPR-Cas13a
    UCA1, MALAT1, TFRC
    mCherry, B4GALNT1, ANXA4, HOTTIP, MALAT1, EGFR, EZH2, HRAS, KRAS, NF2, NFKB1, NRAS, PPARG, RAF1, SMARCA4, STAT3

    vlincRNAs (ID30, ID332, ID-372, ID-604)
    RP11-326A19.4
    Cluc W85X
    Gluc/Cluc luciferases PPIB, KRAS, MALAT1 CXCR4
    RFP
    Trans-ssRNA
    Knock out of long noncoding RNA genes by the deletion of promoters
    Programmable RNA targeting and cleavage
    Targeted knockdown of vlincRNAs
    Activation of long noncoding RNA promoter region inducing genome wide transcriptional changes
    ADAR2 domain fusion with dCas13 for targeted RNA editing through the direct adenosine to inosine deaminase activity of ADAR2
    Engineering of Cas7-11 for RNA targeting and knockdown
    Guided cleavage of ssRNAs by CRISPR-Cas13a effector
    [36,70–76,86]
    TF hinderanceCRISPR-dCas9OCT4, Nanog, Pax6By Steric hinderance of transcription factors by dCas9 targeted binding leading to disturbed TF-DNA interaction[87]
    Genome organizationdCas9-DNMT3A
    dCas9 fused with PYL or ABI proteins (CLOuD9)
    dCas9-CIBN
    miR290, Pou5f1
    β-globin promoter, LCR HS2, OCT4
    Klf4, Zfp462
    promoter
    Targeted methylation of CTCF anchor sites blocking CTCF binding and altering DNA looping and expression of genes located in neighboring loops
    Utilizing CIP for selective formation of de novo chromatin loops dCas9 proteins can be fused to ABI and PYL that can dimerize when the small chemical inducer ABA is introduced to reversibly bring two genomic regions into juxtaposition
    LADL system that drives de novo chromatin looping by the heterodimerization of dCas9-CIBN fusion proteins when blue light is applied using cryptochrome 2 inducible bridging factor
    [56,88,89]

    CIP: Chemically induced proximity; LADL: Light-activated dynamic looping; MCP: MS2 coat protein; ssRNA: Single-stranded RNA; TF: Transcription factor; vlincRNA: Very long intergenic noncoding RNA.

    Mechanisms of epigenome targeting by CRISPR systems

    DNA methylation

    DNA methylation is a major epigenetic repressive chemical modification that influences gene expression. This modification is established by the addition of a methyl group to the fifth carbon position of cytosine to form 5-methylcytosine (5mC). ‘Writer’ proteins, namely DNA methyltransferases (Dnmts), catalyze this process [4], while TET1–3 enzymes control active removal of the methylation modification by adding a hydroxyl group to the methyl group of 5mC to produce a 5hmC modification that can be converted back to bare cytosine by the oxidase activity of TET enzymes [90]. Many epigenome-editing technologies developed for DNA methylation control make use of these ‘writer’ and ‘eraser’ proteins to add or remove chemical modifications to specific target genome loci using dCas9 systems when fused with Dnmts or TET domains. This is demonstrated in previous studies [49,53–61] where Dnmt3a, Dnmt3L and TET1 catalytic domains were fused to CRISPR/dCas9 to create targeted DNA methylation control tools. This has been used along with engineered sgRNA molecules to target and bind genomic loci, recruit enzymatic DNA modification activity, and eventually modulate upregulation or downregulation of gene expression of the target genes. These systems has given us the means to investigate transcriptional changes in response to alterations in localized epigenomic marks. Other research by Pflueger et al. aimed to optimize the specificity of such systems that relies on fusion proteins for DNA methylation modulation [62]. As the understanding of the off-target effects of such systems is lacking, and due to the importance of specificity in the implementation of epigenome-editing tools, a modular dCas9-SunTag DNMT3A dual system was designed. The system consists of a dCas9 protein fused to a SunTag peptide array and a counterpart antibody to a DNMT3A catalytic domain. This research revealed that the dual transfection of dCas9-SunTag and DNMT3A single-chain antibody fusion (αGCN4-D3A) construct leads to more specific and potent methylation of target sites. This is attributed to the epitope activity of the SunTag peptide that recruits multiple (αGCN4-D3A) molecules to the target site. The dual nature of this system also allows for independent variation in the expression of the DNA-targeting module (dC9Sun) and the effector module (GCN4-D3A) [62]. A more recent innovative approach uses a dCas9-based systems with no epigenetic modifying enzymes to control DNA methylation in a less confounding way via steric hindrance. This method of epigenome control is based on the fact that dCas9-gRNA binding to a specific site will block binding and subsequently the activity of DNA methyltransferases in dividing cells. Blocking access of the enzymes to the target site will lead to passive demethylation of the site by dilution as the cells divide. Considering that steric hindrance requires strong, tight binding between the dCas9-gRNA and the target site, off-target effects will be averted [77]. In a seminal paper by Nuñez et al., novel CRISPR-based epigenome engineering tools, ‘CRISPRon’ and ‘CRISPRoff’, were reported. The development of these systems used the dCas9 and catalytic protein fusion approach, where CRISPRoff is composed of ZNF10 KRAB, Dnmt3A (D3A) and Dnmt3L (D3L) protein domains fused to catalytically null dCas9. This system showed the potential to silence the vast majority of human genes robustly and specifically. In addition, it can target and disturb enhancers as well as genes lacking canonical CpG islands. The silencing of genes by CRISPRoff expression is stably maintained throughout cell divisions and stem cell differentiation, which provides a powerful tool to study the mechanisms and roles of epigenetic modifications and inheritance. CRISPRoff also provides a platform for genome-wide screens, multiplexed cell engineering and, in general, a tool to rewrite the human epigenome. Effects of CRISPRoff can be reversed by CRISPRon and can even optimize the gene expression of target genes. CRISPRon is composed of a TETv4 and MCP-transactivator domain (VP64, p65-AD and Rta) fusion proteins that recruit the transcriptional machinery leading to a higher level of expression in target genes [46].

    Histone modifications

    Histone proteins are the basic components of chromatin that are post-translationally modified in a number of ways such as acetylation, methylation, phosphorylation, ubiquitination, and so on [91]. These modifications can regulate and remodel the chromatin structure to up- or down-regulate the transcription of genes by the recruitment of regulatory proteins and complexes [91,92]. Epigenome editing tools have been developed that make it possible to target and ‘tweak’ those chemical modifications to control cellular behaviors. Changes in chromatin state can be achieved by the dCas-based recruitment of effectors to deposit or remove histone acetylation and methylation marks [47,63–66,78–81,93]. Depositing histone acetylation patterns by this method have been a successful approach in the locus-specific transactivation of gene expression [47,63–65,78–81,93]. For instance, acetylation of the histone3 lysine 27 residue is an extensively targeted modifications by CRISPR-based editing systems. This was done by designing a nuclease-null dCas9 protein fused to the catalytic domains of acetyltransferases such as p300 [47,64,65,78–81] to modulate proximal and distal enhancer-regulated genes. These studies showed that dCas9-p300 fusion proteins are effective in transactivating endogenous genes. To analyze the functions of active enhancers based on enhancer-associated chromatin modifications, enCRISPRa and enCRISPRi dual-effector editing systems were developed. The enCRISPRa system combines acetylation writing domain p300 and transcriptional effector domain VP64 for enhancer activation, while the antagonistic enCRISPRi system is composed of the LSD1 lysine demethylase domain along with a KRAB transcriptional repressor domain for maximal perturbation of enhancer functions [80]. CRISPR-Cas9 histone deacetylases (HDACs) protein fusions for transcriptional repression have also been developed, wherein the dCas9-HDAC3 fusion system represses transcription at endogenous promoters [66]. Histone lysine methylation is another critical epigenetic modification that regulates gene expression in a context-dependent manner. Typically, methylation of H3K4, H3K36 and H3K79 residues is associated with an active state, while H3K9, H3K27 and H4K20 methylation is associated with a repressed chromatin state [82,94,95]. Histone methylation dynamics is regulated by methyltransferases, the writers, and demethylases, the erasers. Several studies reported the development of dCas9-based systems to modulate histone lysine methylation marks [63,67–69,83,85,96,97]. The dCas9 fusion proteins were constructed with different methyltransferases that target different histone lysine residues: repressive trimethylation at H3K27Me3 by EZH2 fusion systems [68,69,85,96], transcription-activating H3H4me3 modifications by SMYD3 and PRDM9 fusions [83,97], and H3K4Me3 demethylation by LSD1 domains to perturb enhancers [63,67]. This targeted recruitment of epigenetic effectors not only achieved the respective epigenetic state with corresponding gene-expression patterns but also allowed us to study the crosstalk and interplay between epigenetic modifications [63]. It is pertinent to note that persistent and potent modulation using dCas9 fusion proteins can be better achieved by multiple effector systems [46,68,80]. This approach suggests promising avenues for the optimization of existing dCas9-based epigenome editing tools.

    RNA engineering

    Although ncRNAs do not encode proteins, they are involved in many cellular functions to regulate gene expression. Recently, many classes of regulatory ncRNAs have been discovered due to advances in transcriptome sequencing techniques [98]. One of the classes of ncRNAs is long non-coding RNAs (lncRNA), which are RNA transcripts usually with a length of more than 200 nucleotides. lncRNAs play major roles in both transcriptional and post-transcriptional regulation, genomic imprinting and chromatin remodeling [99–101]. Owing to their roles in cellular processes and human disorders, the functional studies of lncRNAs are of paramount importance to elucidate their diverse roles in gene regulation and cellular functions. While RNAi for ncRNAs has limited success, CRISPR-based systems proved to be powerful in the genetic manipulation of lncRNAs, providing versatile and effective tools for lncRNA studies. The potential of CRISPR systems as a screening tool was demonstrated in numerous studies [86,102–108]. CRISPR-based lncRNA editing approaches vary from targeted deletions or mutations [86,102,103,108], targeting splice sites [104] and CRISPRi/CRISPRa systems [105,107], where transcription is repressed or activated by dCas9 with transcription effector domain fusions such as KRAB and VP64. In addition, the lncRNA targeting toolbox also includes direct targeting of the lncRNA molecules by the RNA binding Cas13 variant to induce knockdowns of lncRNAs without genome alterations [36,70]. Although these tools show efficiency and specificity, they still have limitations and challenges including off-target effects on neighboring gene-expression levels as a result of transcription effector recruitment and unrelated promoter interactions [71].

    As discussed earlier, type VI CRISPR-Cas systems contain RNA targeting Cas13 effectors that can bind and cleave single-stranded RNAs, making them attractive tools for RNA editing and transcriptome modulation. Cas13 proteins contain conserved higher eukaryote and prokaryote nucleotide-binding domains that have nuclease activity for RNA cleavage and degradation of single-stranded RNAs [42,109]. This RNAse activity induces transcript knockdowns in mammalian cells [72–75,110], thus demonstrating transcriptome engineering without permanent alterations to the DNA sequence. Moreover, a catalytically inactive Cas13 protein was engineered that serves as an RNA-targeting platform without the induction of RNA cleavage. This allowed the fusion of RNA editing proteins such as adenosine deaminases acting on RNA (ADARs) that can catalyze the conversion of adenosines to inosines with effects on translation and splicing [111,112], creating an RNA-editing platform with a similar approach as dCas9 fusion effectors [72]. The dCas13b RNA targeting platform has a number of applications including transcript modifications, splicing modification, RNA imaging and transcripts localization within the cell. With this promising expansion of CRISPR epigenome editing tools, further research and consideration is needed to interrogate the toxic effects of targeted RNA degradation. A recent study by Özcan et al. explored the use of class 1 subtype III-E effector Cas7-11 derived from fusion between a Cas11 domain and multiple Cas7 subunits, as a programmable RNA cleavage system. In contrast to RNA knockdown systems based on Cas13 effectors, Cas7-11 RNA knockdown exhibited negligible collateral activity and cell toxicity [76]. This finding could pave the way for the development of new RNA-targeting platforms based on cas7-11 orthologs that are more effective and safer than currently available platforms in the pursuit of RNA-targeting therapeutics.

    Regulatory regions

    Traditionally known as ‘junk DNA’, the noncoding region of our genome plays a pivotal role in the regulation and 3D organization of our genome. These cis-regulatory regions modulate the epigenetic landscape via interactions with histones, sequence-specific DNA-binding transcription factors and other regulatory proteins, thus regulating gene expression at distal sites [113]. Alterations in noncoding regions may lead to the deregulation of cognate genes, even if the regulatory sequences are located far away from the gene itself [114]. Identification and unraveling of functions of noncoding genomic regions will help better understand interactions between gene regulatory programs. CRISPR-based regulatory region annotation screens [80,84,87,115–123] provided a high-throughput assay that enabled us to study the activities of endogenous regulatory elements and identify target genes of regulatory regions. In general, CRISPR-cas9 systems rely on introducing indels to regulatory genomic locations such as transcription factor binding sites and enhancers to examine the effects of loss of function in these regions on target genes [124]. However, such approaches fail to quantify the impact of the regulatory elements on the transcriptional output, along with the fact that these only rely on loss-of-function analysis and do not include gain-of-function studies. To mitigate these limitations, dCas9-KRAB repressor and 9Cas9-p300 activator epigenetic-modifying fusion proteins have been utilized to study both loss- and gain-of-function effects on gene promoters and both proximal and distal genomic enhancer regions [125,126]. This approach has proven to be instrumental in studying gene regulation as it allows us to precisely manipulate regulatory regions and dissect their functions. Furthermore, dCas9 fusion proteins can be combined with other screening approaches such as RNA fluorescence in situ hybridization labeling, high-throughput sequencing and flow cytometry to rigorously quantify target transcript levels and identify candidate regulatory elements [127].

    3D genome organization is another level of regulation that determines how the genome is hierarchically packaged inside the nucleus. This entails organization of the chromatin as chromosome territories, euchromatin and heterochromatin compartments, topologically associating domains, and chromatin loops. 3D chromatin architecture is a key determinant of gene expression patterns and is altered in response to environmental stress and developmental cues [128,129]. CRISPR-based epigenome regulation has been employed to delineate the nature of chromatin looping. One system employed by Morgan et al. is CLOuD9, a dCas9-based approach to forcibly introduce loops in the chromatin structure, bringing together any two genomic regions to gain insight into their interactions and effects on gene expression. CLOuD9 consists of dCas9 fusion with components of the plant ABA signaling pathway. dCas9 fusion proteins are targeted to each genomic loci by sgRNAs; upon the addition of phytohormone S-(+)-abscisic acid (ABA), dimerization occurs between the chemical-induced proximity proteins. This results in interaction and juxtaposition of the two chromosomal loci [88]. Similarly, Kim et al. developed an optogenetic-inducible CRISPR-based tool called the light-activated dynamic looping system to direct a stretch enhancer to the Zfp462 promoter using CRISPR gRNAs and light-induced heterodimerized cryptochrome 2 and a dCas9-CIBN fusion protein [89]. Such platforms offer inducible, reversible and broadly applicable tools for 3D chromatin manipulation that likely will prove to be of significance in studying abnormal chromatin effects in human disorders. Figure 2 summarizes the tools in use for epigenome and RNA editing.

    Figure 2. Epigenome and RNA editing via CRISPR.

    (A) Cas protein (light blue) binds DNA target sequence (orange) via guide RNA (orange line) near protospacer adjacent motif (PAM) (blue). This leads to cleavage of the target sequence due to nuclease activity of the Cas protein, producing double-stranded breaks that will be mend by the endogenous DNA repair machinery. (B) dCas lacks the ability to induce breaks in target sequences. However, can still bind to target sequences via gRNAs and add/remove specific epigenetic marks or modulate chromosome looping. (C) Cas13, guided by gRNA, targets the RNA (yellow) molecule for cleavage. (D) However, dCas13 juxtaposed with ADAR2 carries out RNA editing by converting A to I.

    ADAR: Adenosine deaminase acting on RNA; dCas9: Nuclease-defective Cas9; gRNA: Guide RNA.

    Applications of CRISPR epigenome editing

    As epigenetic reprogramming underlies a number of developmental and environmental disorders [130,131], manipulation of the epigenome offers a multitude of biological applications. CRISPR-mediated epigenome editing can be applied to address challenges in the fields of drug screening, disease modelling and cell fate engineering. Soon after conception, CRISPR epigenome editing has been used to investigate both basic biological questions and clinical applications in various fields. One exciting area of basic exploration includes identifying and exploring the functions of noncoding regulatory regions of the genome. Enhancers are one such region with characteristics epigenetic marks, H3K4me1/2 and H3K27ac [132–134]. CRISPR/dCas9-based enhancer-targeting epigenetic editing systems have been developed that provide opportunities to identify corresponding genes regulated by specific enhancers as well as interrogating enhancer function in native biological contexts [80,127]. In addition to these basic biological discoveries, the expansion in the CRISPR epigenome engineering toolbox served as a catalyst for major developments in stem cell research, biotechnology, regenerative medicine and basic research (summarized in Table 3). Here we overview their applications in the field of cancer biology and regenerative medicine.

    Table 3. Applications of CRISPR mediated epigenome editing in cancer research and regenerative medicine.
    ApplicationTargetMechanismRef.
    Cancer therapeuticsGRN in Hep3B hepatoma cells
    SARI promoter in colon cancer tissue
    DKK3 promoter in PCa cells
    ΔNp63 in in lung and esophageal SCCs
    Two class II TSGs, MASPIN, REPRIMO in H157 lung cancer cells and AGS gastric cancer cells
    MDM2 in osteosarcoma cells
    KLF4 in UBC
    Rnd3 in MM RPMI 8226 and JJN3 cell lines
    Targeting of a dCas9 fusion system with DNMT3a, EZH2 and KRAB epi-suppressors to the GRN promoter, leading to knockdown and reduction of tumor formation
    Targeted demethylation of the SARI promoter by a dCas9 TET1 fusion system, leading to restoring SARI antitumor function
    Targeted demethylation of the DKK3 promoter by a dCas9-VPR fusion system, leading to induction of DKK3 tumor suppressive roles
    CRISPR interference system (dCas9-KRAB) targeted to the transcription start site of ΔNp63, leading its suppression and induced apoptosis of cancer cells
    CRISPR/dCas9 fused with a number of effector domains (VP64, p300, VPR, SAM complex) for dormant tumor suppressor gene reactivation to inhibit cell proliferation
    Control of the MDM2 proto-oncogene by dCas9-KRAB in osteosarcoma cells, leading to effective inhibition of tumor growth
    Upregulation of KLF4 expression by dCas9-VP64 (CRISPRon), leading to the inhibition of tumorigenesis
    Silencing of Rnd3 in multiple myeloma cells by dCas9-KRAB
    [96,135–145]
    Regenerative medicineUCP1 in adipose-derived stem cells (ADSCs)
    Pparγ2, Prdm16, Zfp423, Ucp1 in adipocytes
    Ascl1, Lmx1a, Nr4a2 (ALN) or Ascl1, Lmx1a, NeuroD1 transcription factors
    FMR1 CGG expansion in induced pluripotent stem cells
    Brn2, Ascl1, Myt1l in fibroblasts
    Upregulation of UCP1 expression in engineered adipocytes using a CRISPR activation system (CRISPRa) composed of dCas9 and activation domains MCP-p65-HSF1
    Upregulation of genes involved in adipocyte differentiation and function by the CRIPSRa SAM system for adipocyte engineering
    Reprogramming of striatal astrocytes into mature neurons by the CRISPRa system dCas9-VP64/SAM for voluntary motor behavior rescue in Parkinson's disease
    Targeted demethylation and reactivation of FMR1 by dCas9-TET1
    Reprogramming mouse embryonic fibroblasts to induced neuronal cells utilizing dCas9-VP64 to activate endogenous neurogenic genes Brn2, Ascl1, Myt1l
    [146–154]

    GRN: Granulin; MM: Multiple myeloma; PCa: Prostate cancer; SAM: Synergistic activation mediator; SCC: Squamous cell carcinoma; UBC: Urothelial bladder cancer.

    Cancer therapeutics

    Epigenetic anomalies contribute to various human diseases, imprinting disorders [135,155,156] and cancer [136,157] in particular due to their critical roles in growth related pathways. The development of CRISPR-based systems has ushered in a new era of cancer therapy, with the potential to precisely target and correct epigenetic alterations that drive tumor growth. Epimutations of tumor-suppressor genes and oncogenes can be targeted via dCas9 fusion proteins with epigenomic ‘writers’ and ‘erasers’ of epigenetic marks such as DNA and histone methylation (Dnmts, EZH2, TET) or transcriptional effectors (VP64, p300, KRAB) to modulate the expression of genes of interest. Studies of dCas9 targeting platforms demonstrated promising results in the suppression of cell proliferation and tumor growth in different types of cancers both in vivo and in vitro following epigenetic targeting. Hypermethylation is found to contribute to carcinogenesis through the silencing of tumor suppressor genes [137], leading gaining malignancy hallmarks. For example, the inactivation of SMARCA2, a key component of the SWI/SNF complex with vital roles in chromatin remodeling gene regulation by promoter hypermethylation, leads to loss of control over cell growth and a significantly reduced survival rate in lung cancer patients. This was experimentally verified by hypermethylation-mediated downregulation of SMARCA2 by targeting the promoter CpG with the dCas9-DNMT3A fusion protein [138]. Thus, the role of epigenome in diseases can be unravelled by CRISPR-mediated modeling of epigenetic aberrations. Several studies have demonstrated the applications of dCas9 fusions with epigenetic effectors that act on DNA methylation marks to target tumor-suppressor genes and oncogenes to modulate their functions for tumor growth inhibition [96,139,140]. For example, BRCA1, a tumor-suppressor gene silenced by DNA methylation, was demethylated at the promoter by dCas9–TET1 fusion, leading to rescued expression of BRCA1 and inhibition of cell proliferation in a cancer cell line [54]. Besides epigenetic effectors, dCas9 systems can be fused to transcription effector domains to target tumor-suppressor genes and oncogenes [141–144,146,147]. However, transcriptional modulators affect gene expression in a broad way, while epigenetic effectors target specific epigenetic changes. This specificity is a key feature of precision therapy, which aims to treat cancer by targeting the specific genetic changes that drive it [146]. Besides therapeutic applications, dCas9 platforms can be used to induce the expression of endogenous oncogenes to generate in vitro models for preclinical studies of cancer therapeutics [145,158].

    Regenerative medicine

    Acquiring the ability to reinstate the function of a lost/damaged organ has been the target of stem cell biologists. The seminal work of Takahashi and Yamanaka identified unique factors that can reprogram differentiated cells to a pluripotent state [159]. These cells, termed induced pluripotent stem cells (iPSCs), are valuable tools for both basic research and clinical applications [148,149]. In basic research, iPSCs can be used to model and study the development of various diseases, which can help us to better understand these processes and screen and develop new drugs. In clinical settings, iPSCs have been used to generate patient-specific cell types for transplantation [150]. Multiple versions of CRISPR have been applied for endogenous gene manipulation to achieve reprogramming [151–153,160]. A dCas9-synergistic activation mediator (SAM)-based screen not only successfully reprogram primed murine epiblast stem cells to a pluripotent embryonic stem cell state but also identified novel molecules implicated in reprogramming that were previously unknown regulators of pluripotency [161].

    Cellular reprogramming, differentiation and transdifferentiation of cells into a required cell type are complex processes that involve the coordinated expression of many genes. By activating or inactivating the required genes, it is possible to direct cells for guided differentiation or transdifferentiation. The ability to activate multiple genes simultaneously is a challenging task but essential for cellular reprogramming. Robustness of CRISPR-dCas9 has been used to achieve this task. For instance, the dCas9-VP64 activator and dCas9-SAM and dCas9-SunTag systems have been successfully used to simultaneously activate several endogenous genes in human PSCs and mesenchymal stem cells with robustness and specificity [154]. Importantly, the systems were effective for single and multiplexed gene activation. Direct reprogramming of mouse fibroblasts to neuronal cells was achieved via activation of neurogenic factors Brn2, Ascl1 and Myt1l (BAM factors) genes by a dCas9-VP64 activation system. This study also suggests that targeting endogenous genes is a more effective approach to achieve sustained cell reprogramming compared with forced expression of transgenes [162]. DNA methylation modulation by dCas9–TET1 fusions showed potency in ameliorating repressive hypermethylation of the causative CGG expansion mutation present in the FMR1 gene in iPSCs derived from fragile X syndrome patients. These adjustments in DNA methylation patterns were shown to be stable and restore the functionality of the iPSC-derived neurons [160]. All these studies demonstrate the power of using CRISPR-mediated epigenome editing in understanding molecular mechanisms that mediate complex cellular processes in regenerative medicine.

    Limitations of current methods

    Although CRISPR systems show high potential as a therapeutic epigenome-editing approach, many challenges persist and make it challenging to implement the technology in clinical trials. First, CRISPR systems are generally prone to off-target effects, which can be a result of using suboptimal gRNA molecules or might be caused by the expression levels of effector domains or dCas proteins, as with increasing expression levels the off-target effects also increase. This is besides the influence of their intrinsic binding specificity and exposure time (time span of expression in the cell) and can be circumvented by using tunable dCas fusions and inducible systems for better control of these parameters. Second, persistence of the therapeutic effects poses another challenge, as epigenetic marks in growing cells may fade away during successive cell divisions by dilution. Therefore, it is important to carry out epigenetic editing in a way that mimics the endurance of natural epigenomic stability. This can be potentially done by creating systems that can actively induce endogenous maintenance mechanisms, to ensure ‘self-sustainability’ of the process. Last, one of the major challenges that limits the application of epigenome editing systems is delivery to target cells. The ideal vehicle is a system that is safe, specific and stable, and with high loading capacity. Although lentiviral systems are most commonly used in clinical settings, the issues of immunogenicity, limited loading capacity and genomic integrations remain major concerns. Recently, delivery systems have been developed that utilize nanoparticles as carriers for CRISPR systems. Nanocarrier systems are an attractive approach as they exhibit high loading capacity and stability. However, more studies are required to characterize and improve their safety [163–165].

    Conclusion

    CRISPR based epigenome editing has rapidly evolved in relatively short period of time. The development of nuclease-null mutant, dCas9, has led to expansion of additional tools that have been instrumental in our understanding of epigenetic mechanisms in gene regulation. Due to these versatile tools and ease of targeting, this system has become the method of choice for epigenetic research. In addition to deciphering the role of various DNA elements, the CRISPR system has also been harnessed to modify RNA molecules. These tools have been successfully applied both in basic and clinical research. It is anticipated that the methodology will continue to evolve in future, and with the development of efficient tools for safe and targeted delivery, epigenome editing will have broader clinical applications in cancer therapeutics, regenerative medicine and other diseases.

    Future perspective

    It has been just 10 years after the first use of CRISPR in eukaryotic system, and the developments in the field have been astounding. Two developments have been progressing in parallel; the identification of new classes of the CRISPR system in widespread organisms, and remodeling of existing classes for diverse uses. Soon after its use in eukaryotes, CRISPR-Cas9 was repurposed for epigenome editing and a number of tools have been developed for this purpose. Epigenome editing has widespread applications ranging from basic biological research to clinical applications. One of the fundamental questions epigenome editing addressed is that it delineated the causal relationships between epigenetic marks and gene expression. The systems developed have the potential to identify novel therapeutic targets. RNA-targeting CRISPR-based platforms have been developed for use in transcriptome engineering and diagnostics.

    Despite all these strides, significant work is needed to improve the specificity of the developed tools. The collateral effect of Cas13, for instance, is a major challenge that has to be addressed, which is being done with the identification of more CRISPR platforms harboring RNA-targeting activity. Another major concern is the utility of these developed tools in vivo. For the in vivo expression of dCas9 fusion proteins in transgenic mice, two conditional transgenic mouse lines were recently developed: Rosa26:LSL-dCas9-p300 for gene activation and Rosa26:LSL-dCas9-KRAB for gene repression [166]. Adeno-associated viruses are the most commonly used vector systems for the in vivo delivery of CRISPR-mediated editing systems [161,167,168] that have been used for disease modeling and treatment of various conditions. Cell-penetrating peptides [169], gold nanoparticles [170], lipid nanoparticles [171] and microinjections [172] are other available methods that have been used for in vivo disease models. However, further development of safe and effective methods to deliver epigenome-editing components still presents a major goal for in vivo genome-editing therapy. In addition, reducing off-targets and ensuring efficient delivery have to be addressed before making a transition from bench to clinic.

    While these limitations currently exist, the efforts to identify novel and efficient genome-editing tools continue. In an effort to identify a eukaryotic counterpart of CRISPR, the Zhang laboratory recently identified the ‘Fanzor’ protein, which is a eukaryotic programmable RNA-guided endonuclease that makes more precise changes in human cells [173]. This not only complements the existing tools, but also points to the possibility of even more diverse genome-editing tools in nature. The identification of new genome-editing tools and their modified versions for epigenome editing hold great promise for future research and therapeutics.

    Executive summary
    • The CRISPR system is a breakthrough gene-editing technology that typically employs a guide RNA to target a specific DNA region.

    • Six different forms of CRISPR-Cas (I–VI) can be distinguished, and each type includes a unique Cas nuclease protein. Based on the makeup of the Cas effectors and CRISPR RNA ribonucleoprotein complexes, the six distinct types have been further classified into two classes.

    • Recently, CRISPR-Cas9 technology was modified for epigenome editing by creating a ‘dead’ Cas9 protein (dCas9) or ‘nuclease-null’, which enables us to take advantage of its DNA-binding capabilities to target specific loci for epigenetic alterations without cleavage.

    Mechanisms of epigenome targeting by CRISPR systems

    • A significant epigenetic constrictive chemical change that affects gene expression is DNA methylation. To make methylation more effective and specific, dCas9-SunTag and DNMT3A single-chain antibody fusion (GCN4-D3A) constructs can be integrated.

    • Another mechanism by which CRISPR can be used for gene editing is through histone modifications. Histone proteins can be altered post-translationally in a variety of ways, including acetylation, methylation, phosphorylation, ubiquitination, and so on.

    • The development of these epigenome editing technologies has made it possible to make targeted chemical modifications to control cellular behaviors.

    • Several noncoding RNAs have been investigated for their role in epigenetic modification, one important one being long noncoding RNAsn (lncRNAs). lncRNAs play a variety of roles in gene regulation and cellular function, as well as a significant role in cellular processes and diseases in humans. CRISPR-based systems have demonstrated effectiveness in lncRNA genetic manipulation, proving to be a flexible and useful tool for lncRNA investigations.

    • CRISPR systems that have been developed to target RNA include the Cas13 effector systems and Cas7–11 RNA systems. These systems allow for RNA editing and transcriptome modification even without the induction of RNA cleavage.

    Regulatory regions

    • Using CRISPR-based regulatory area annotation screens, we were able to identify and decipher the roles of noncoding genomic regions.

    • dCas9 fusion proteins can be used in conjunction with other screening techniques to study both loss- and gain-of-function effects on gene promoters.

    • The nature of chromatin looping has been defined using CRISPR-based epigenome regulation. The use of 3D chromatin manipulation systems such as the CLOuD9 and light-activated dynamic looping systems is pivotal for understanding the role of aberrant chromatin looping in human illnesses.

    Applications of CRISPR epigenome editing

    • Oncogenes and tumor suppressor genes can undergo epimutation, which can be targeted by dCas9 fusion proteins. The normal expression of tumor suppressors may be rescued.

    • Cells can be directed for pluripotency using a dCas9-synergistic activation mediator-based screen.

    • The CRISPR-dCas9 system has been utilized to simultaneously activate many genes that have important implications in regenerative medicine.

    • Numerous fields, including fundamental biological research and therapeutic applications, use CRISPR-based epigenome editing.

    • Further work on CRISPR-based system for epigenome editing holds great significance for the future.

    Financial disclosure

    The authors have no financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending, or royalties.

    Competing interests disclosure

    The authors have no competing interests or relevant affiliations with any organization or entity with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending, or royalties.

    Writing disclosure

    No writing assistance was utilized in the production of this manuscript.

    Open access

    This work is licensed under the Creative Commons Attribution 4.0 License. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/

    Papers of special note have been highlighted as: • of interest; •• of considerable interest

    References

    • 1. Waddington CH. The epigenotype. 1942. Int. J. Epidemiol. 41(1), 10–13 (2012).
    • 2. Li Y, Chen X, Lu C. The interplay between DNA and histone methylation: molecular mechanisms and disease implications. EMBO Rep. 22(5), e51803 (2021).
    • 3. Zhang Y, Sun Z, Jia J et al. Overview of histone modification. Adv. Exp. Med. Biol. 1283, 1–16 (2021).
    • 4. Moore LD, Le T, Fan G. DNA methylation and its basic function. Neuropsychopharmacology 38(1), 23–38 (2013).
    • 5. Klemm SL, Shipony Z, Greenleaf WJ. Chromatin accessibility and the regulatory epigenome. Nat. Rev. Genet. 20(4), 207–220 (2019).
    • 6. Bonev B, Cavalli G. Organization and function of the 3D genome. Nat. Rev. Genet. 17(12), 772 (2016).
    • 7. Morris KV. The emerging role of RNA in the regulation of gene transcription in human cells. Semin. Cell Dev. Biol. 22(4), 351–358 (2011).
    • 8. Morf J, Basu S, Amaral PP. RNA, genome output and input. Front. Genet. 11, 589413 (2020).
    • 9. de Mendoza A, Sebe-Pedros A. Origin and evolution of eukaryotic transcription factors. Curr. Opin. Genet. Dev. 58–59, 25–32 (2019).
    • 10. Barrera LO, Ren B. The transcriptional regulatory code of eukaryotic cells – insights from genome-wide analysis of chromatin organization and transcription factor binding. Curr. Opin. Cell Biol. 18(3), 291–298 (2006).
    • 11. Li H, Cui D, Wu S et al. Epigenetic regulation of gene expression in epithelial stem cells fate. Curr. Stem Cell Res. Ther. 13(1), 46–51 (2018).
    • 12. Namihira M, Kohyama J, Abematsu M, Nakashima K. Epigenetic mechanisms regulating fate specification of neural stem cells. Philos Trans. R. Soc. Lond. B. Biol. Sci. 363(1500), 2099–2109 (2008).
    • 13. Wu H, Sun YE. Epigenetic regulation of stem cell differentiation. Pediatr. Res. 59(4 Pt 2), 21R–25R (2006).
    • 14. Jasencakova Z, Groth A. Restoring chromatin after replication: how new and old histone marks come together. Semin. Cell Dev. Biol. 21(2), 231–237 (2010).
    • 15. Reveron-Gomez N, Gonzalez-Aguilera C, Stewart-Morgan KR et al. Accurate recycling of parental histones reproduces the histone modification landscape during DNA replication. Mol. Cell 72(2), 239–249.e235 (2018).
    • 16. Zhang L, Lu Q, Chang C. Epigenetics in health and disease. Adv. Exp. Med. Biol. 1253, 3–55 (2020).
    • 17. Wu Y, Sarkissyan M, Vadgama JV. Epigenetics in breast and prostate cancer. Methods Mol. Biol. 1238, 425–466 (2015).
    • 18. Garcia-Martinez L, Zhang Y, Nakata Y, Chan HL, Morey L. Epigenetic mechanisms in breast cancer therapy and resistance. Nat. Commun. 12(1), 1786 (2021).
    • 19. Nebbioso A, Tambaro FP, Dell'Aversana C, Altucci L. Cancer epigenetics: moving forward. PLOS Genet. 14(6), e1007362 (2018).
    • 20. Moore-Morris T, van Vliet PP, Andelfinger G, Puceat M. Role of epigenetics in cardiac development and congenital diseases. Physiol. Rev. 98(4), 2453–2475 (2018).
    • 21. Gomes CPC, Schroen B, Kuster GM et al. Regulatory RNAs in heart failure. Circulation 141(4), 313–328 (2020).
    • 22. Cao J, Wu Q, Huang Y, Wang L, Su Z, Ye H. The role of DNA methylation in syndromic and non-syndromic congenital heart disease. Clin. Epigenetics 13(1), 93 (2021).
    • 23. Rohde K, Keller M, la Cour Poulsen L, Bluher M, Kovacs P, Bottcher Y. Genetics and epigenetics in obesity. Metabolism 92, 37–50 (2019).
    • 24. Gao W, Liu JL, Lu X, Yang Q. Epigenetic regulation of energy metabolism in obesity. J. Mol. Cell Biol. 13(7), 480–499 (2021).
    • 25. Lardenoije R, Iatrou A, Kenis G et al. The epigenetics of aging and neurodegeneration. Prog. Neurobiol. 131, 21–64 (2015).
    • 26. Lardenoije R, Pishva E, Lunnon K, van den Hove DL. Neuroepigenetics of aging and age-related neurodegenerative disorders. Prog. Mol. Biol. Transl. Sci. 158, 49–82 (2018).
    • 27. Majchrzak-Celinska A, Baer-Dubowska W. Pharmacoepigenetics: an element of personalized therapy? Expert Opin. Drug Metab. Toxicol. 13(4), 387–398 (2017).
    • 28. Garcia-Gimenez JL, Sanchis-Gomar F, Lippi G et al. Epigenetic biomarkers: a new perspective in laboratory diagnostics. Clin. Chim Acta 413(19–20), 1576–1582 (2012).
    • 29. Miranda Furtado CL, Dos Santos Luciano MC, Da Silva Santos R, Furtado GP, Moraes MO, Pessoa C. Epidrugs: targeting epigenetic marks in cancer treatment. Epigenetics 14(12), 1164–1176 (2019).
    • 30. Halabian R, Valizadeh A, Ahmadi A, Saeedi P, Azimzadeh Jamalkandi S, Alivand MR. Laboratory methods to decipher epigenetic signatures: a comparative review. Cell Mol. Biol. Lett. 26(1), 46 (2021).
    • 31. Waryah CB, Moses C, Arooj M, Blancafort P. Zinc fingers, TALEs, and CRISPR systems: a comparison of tools for epigenome editing. Methods Mol. Biol. 1767, 19–63 (2018).
    • 32. Mali P, Yang L, Esvelt KM et al. RNA-guided human genome engineering via Cas9. Science 339(6121), 823–826 (2013). •• One of the earliest ones to show CRISPR-Cas9 harnessed for genome editing in eukaryotes.
    • 33. Makarova KS, Wolf YI, Alkhnbashi OS et al. An updated evolutionary classification of CRISPR-Cas systems. Nat. Rev. Microbiol. 13(11), 722–736 (2015).
    • 34. Koonin EV, Makarova KS, Zhang F. Diversity, classification and evolution of CRISPR-Cas systems. Curr. Opin. Microbiol. 37, 67–78 (2017).
    • 35. Liu TY, Iavarone AT, Doudna JA. RNA and DNA targeting by a reconstituted Thermus thermophilus type III-A CRISPR-Cas system. PLOS ONE 12(1), e0170552 (2017).
    • 36. Konermann S, Lotfy P, Brideau NJ, Oki J, Shokhirev MN, Hsu PD. Transcriptome engineering with RNA-targeting type VI-D CRISPR effectors. Cell 173(3), 665–676.e614 (2018). • Along with Cox et al. and Abudayyeh et al. [101,102], showed that Cas13 can be used to efficiently target RNA molecules.
    • 37. Perculija V, Lin J, Zhang B, Ouyang S. Functional features and current applications of the RNA-targeting type VI CRISPR-Cas systems. Adv. Sci. (Weinh.) 8(13), 2004685 (2021).
    • 38. Xu X, Qi LS. A CRISPR-dCas toolbox for genetic engineering and synthetic biology. J. Mol. Biol. 431(1), 34–47 (2019).
    • 39. Wang Q, Liu Y, Han C et al. Efficient RNA virus targeting via CRISPR/CasRx in fish. J. Virol. 95(19), e0046121 (2021).
    • 40. Gupta R, Ghosh A, Chakravarti R et al. Cas13d: a new molecular scissor for transcriptome engineering. Front. Cell. Dev. Biol. 10, 866800 (2022).
    • 41. Yan WX, Chong S, Zhang H et al. Cas13d is a compact RNA-targeting type VI CRISPR effector positively modulated by a WYL-domain-containing accessory protein. Mol. Cell. 70(2), 327–339.e325 (2018).
    • 42. O'Connell MR. Molecular mechanisms of RNA targeting by Cas13-containing type VI CRISPR-Cas systems. J. Mol. Biol. 431(1), 66–87 (2019).
    • 43. Qi LS, Larson MH, Gilbert LA et al. Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell 152(5), 1173–1183 (2013). •• Reported catalytically dead Cas9 lacking endonuclease activity that, along with guide RNA, generates a DNA recognition complex.
    • 44. Enriquez P. CRISPR-mediated epigenome editing. Yale J. Biol. Med. 89(4), 471–486 (2016).
    • 45. Cong L, Ran FA, Cox D et al. Multiplex genome engineering using CRISPR/Cas systems. Science 339(6121), 819–823 (2013).
    • 46. Nunez JK, Chen J, Pommier GC et al. Genome-wide programmable transcriptional memory by CRISPR-based epigenome editing. Cell 184(9), 2503–2519.e2517 (2021). • Describes a method that uses CRISPRoff as a single fusion protein that programs heritable epigenetic memory.
    • 47. Hilton IB, D'Ippolito AM, Vockley CM et al. Epigenome editing by a CRISPR-Cas9-based acetyltransferase activates genes from promoters and enhancers. Nat. Biotechnol. 33(5), 510–517 (2015).
    • 48. Chen E, Lin-Shiao E, Trinidad M, Saffari Doost M, Colognori D, Doudna JA. Decorating chromatin for enhanced genome editing using CRISPR-Cas9. Proc. Natl Acad. Sci. USA 119(49), e2204259119 (2022).
    • 49. Vojta A, Dobrinic P, Tadic V et al. Repurposing the CRISPR-Cas9 system for targeted DNA methylation. Nucleic Acids Res. 44(12), 5615–5628 (2016).
    • 50. Xiong T, Meister GE, Workman RE et al. Targeted DNA methylation in human cells using engineered dCas9-methyltransferases. Sci. Rep. 7(1), 6732 (2017).
    • 51. Yang J, Meng X, Pan J et al. CRISPR/Cas9-mediated noncoding RNA editing in human cancers. RNA Biol. 15(1), 35–43 (2018).
    • 52. Phelan JD, Staudt LM. CRISPR-based technology to silence the expression of IncRNAs. Proc. Natl Acad. Sci. USA 117(15), 8225–8227 (2020).
    • 53. Morita S, Noguchi H, Horii T et al. Targeted DNA demethylation in vivo using dCas9-peptide repeat and scFv-TET1 catalytic domain fusions. Nat. Biotechnol. 34(10), 1060–1065 (2016).
    • 54. Choudhury SR, Cui Y, Lubecka K, Stefanska B, Irudayaraj J. CRISPR-dCas9 mediated TET1 targeting for selective DNA demethylation at BRCA1 promoter. Oncotarget 7(29), 46545–46556 (2016).
    • 55. Xu X, Tao Y, Gao X et al. A CRISPR-based approach for targeted DNA demethylation. Cell. Discov. 2, 16009 (2016).
    • 56. Liu XS, Wu H, Ji X et al. Editing DNA methylation in the mammalian genome. Cell 167(1), 233–247.e217 (2016).
    • 57. McDonald JI, Celik H, Rois LE et al. Reprogrammable CRISPR/Cas9-based system for inducing site-specific DNA methylation. Biol Open. 5(6), 866–874 (2016).
    • 58. Huang YH, Su J, Lei Y et al. DNA epigenome editing using CRISPR-Cas SunTag-directed DNMT3A. Genome Biol. 18(1), 176 (2017).
    • 59. Stepper P, Kungulovski G, Jurkowska RZ et al. Efficient targeted DNA methylation with chimeric dCas9-Dnmt3a-Dnmt3L methyltransferase. Nucleic Acids Res. 45(4), 1703–1713 (2017).
    • 60. Josipovic G, Tadic V, Klasic M et al. Antagonistic and synergistic epigenetic modulation using orthologous CRISPR/dCas9-based modular system. Nucleic Acids Res. 47(18), 9637–9657 (2019).
    • 61. Nguyen TV, Lister R. Genomic targeting of TET activity for targeted demethylation using CRISPR/Cas9. Methods Mol. Biol. 2272, 181–194 (2021).
    • 62. Pflueger C, Tan D, Swain T et al. A modular dCas9-SunTag DNMT3A epigenome editing system overcomes pervasive off-target activity of direct fusion dCas9-DNMT3A constructs. Genome Res. 28(8), 1193–1206 (2018).
    • 63. Zhao W, Xu Y, Wang Y et al. Investigating crosstalk between H3K27 acetylation and H3K4 trimethylation in CRISPR/dCas-based epigenome editing and gene activation. Sci. Rep. 11(1), 15912 (2021).
    • 64. Kuscu C, Mammadov R, Czikora A et al. Temporal and spatial epigenome editing allows precise gene regulation in mammalian cells. J. Mol. Biol. 431(1), 111–121 (2019).
    • 65. Shrimp JH, Grose C, Widmeyer SRT, Thorpe AL, Jadhav A, Meier JL. Chemical control of a CRISPR-Cas9 acetyltransferase. ACS Chem. Biol. 13(2), 455–460 (2018).
    • 66. Kwon DY, Zhao YT, Lamonica JM, Zhou Z. Locus-specific histone deacetylation using a synthetic CRISPR-Cas9-based HDAC. Nat. Commun. 8, 15315 (2017).
    • 67. Kearns NA, Pham H, Tabak B et al. Functional annotation of native enhancers with a Cas9-histone demethylase fusion. Nat. Methods 12(5), 401–403 (2015).
    • 68. O'Geen H, Ren C, Nicolet CM et al. dCas9-based epigenome editing suggests acquisition of histone methylation is not sufficient for target gene repression. Nucleic Acids Res. 45(17), 9901–9916 (2017).
    • 69. Chen X, Wei M, Liu X et al. Construction and validation of the CRISPR/dCas9-EZH2 system for targeted H3K27Me3 modification. Biochem. Biophys. Res. Commun. 511(2), 246–252 (2019).
    • 70. Xu D, Cai Y, Tang L et al. A CRISPR/Cas13-based approach demonstrates biological relevance of vlinc class of long non-coding RNAs in anticancer drug response. Sci. Rep. 10(1), 1794 (2020).
    • 71. Soubeyrand S, Lau P, Peters V, McPherson R. Off-target effects of CRISPRa on interleukin-6 expression. PLoS One 14(10), e0224113 (2019).
    • 72. Cox DBT, Gootenberg JS, Abudayyeh OO et al. RNA editing with CRISPR-Cas13. Science 358(6366), 1019–1027 (2017). • One of the seminal papers that showed that Cas 13 can be used to efficiently target RNA molecules.
    • 73. Abudayyeh OO, Gootenberg JS, Essletzbichler P et al. RNA targeting with CRISPR-Cas13. Nature 550(7675), 280–284 (2017). • One of the earliest papers that showed that Cas 13 can be used to efficiently target RNA molecules.
    • 74. Abudayyeh OO, Gootenberg JS, Konermann S et al. C2c2 is a single-component programmable RNA-guided RNA-targeting CRISPR effector. Science 353(6299), aaf5573 (2016).
    • 75. East-Seletsky A, O'Connell MR, Burstein D, Knott GJ, Doudna JA. RNA targeting by functionally orthogonal type VI-A CRISPR-Cas enzymes. Mol. Cell. 66(3), 373–383.e373 (2017).
    • 76. Ozcan A, Krajeski R, Ioannidi E et al. Programmable RNA targeting with the single-protein CRISPR effector Cas7-11. Nature 597(7878), 720–725 (2021).
    • 77. Sapozhnikov DM, Szyf M. Unraveling the functional role of DNA demethylation at specific promoters by targeted steric blockage of DNA methyltransferase with CRISPR/dCas9. Nat. Commun. 12(1), 5711 (2021).
    • 78. Gao D, Liang FS. Chemical inducible dCas9-guided editing of H3K27 acetylation in mammalian cells. Methods Mol. Biol. 1767, 429–445 (2018).
    • 79. Zhang X, Wang W, Shan L et al. Gene activation in human cells using CRISPR/Cpf1-p300 and CRISPR/Cpf1-SunTag systems. Protein Cell 9(4), 380–383 (2018).
    • 80. Li K, Liu Y, Cao H et al. Interrogation of enhancer function by enhancer-targeting CRISPR epigenetic editing. Nat. Commun. 11(1), 485 (2020).
    • 81. Chen T, Gao D, Zhang R et al. Chemically controlled epigenome editing through an inducible dCas9 system. J. Am. Chem. Soc. 139(33), 11337–11340 (2017).
    • 82. Hyun K, Jeon J, Park K, Kim J. Writing, erasing and reading histone lysine methylations. Exp. Mol. Med. 49(4), e324 (2017).
    • 83. Cano-Rodriguez D, Gjaltema RA, Jilderda LJ et al. Writing of H3K4Me3 overcomes epigenetic silencing in a sustained but context-dependent manner. Nat. Commun. 7, 12284 (2016).
    • 84. Thakore PI, D'Ippolito AM, Song L et al. Highly specific epigenome editing by CRISPR-Cas9 repressors for silencing of distal regulatory elements. Nat. Methods 12(12), 1143–1149 (2015).
    • 85. Fukushima HS, Takeda H, Nakamura R. Targeted in vivo epigenome editing of H3K27me3. Epigenetics Chromatin 12(1), 17 (2019).
    • 86. Aparicio-Prat E, Arnan C, Sala I, Bosch N, Guigo R, Johnson R. DECKO: single-oligo, dual-CRISPR deletion of genomic elements including long non-coding RNAs. BMC Genomics 16, 846 (2015).
    • 87. Shariati SA, Dominguez A, Xie S, Wernig M, Qi LS, Skotheim JM. Reversible disruption of specific transcription factor-DNA interactions using CRISPR/Cas9. Mol. Cell. 74(3), 622–633.e624 (2019).
    • 88. Morgan SL, Mariano NC, Bermudez A et al. Manipulation of nuclear architecture through CRISPR-mediated chromosomal looping. Nat. Commun. 8, 15993 (2017).
    • 89. Kim JH, Rege M, Valeri J et al. LADL: light-activated dynamic looping for endogenous gene expression control. Nat. Methods 16(7), 633–639 (2019).
    • 90. Kohli RM, Zhang Y. TET enzymes, TDG and the dynamics of DNA demethylation. Nature 502(7472), 472–479 (2013).
    • 91. Bannister AJ, Kouzarides T. Regulation of chromatin by histone modifications. Cell Res. 21(3), 381–395 (2011).
    • 92. Stillman B. Histone modifications: insights into their influence on gene expression. Cell 175(1), 6–9 (2018).
    • 93. Chiarella AM, Butler KV, Gryder BE et al. Dose-dependent activation of gene expression is achieved using CRISPR and small molecules that recruit endogenous chromatin machinery. Nat. Biotechnol. 38(1), 50–55 (2020).
    • 94. Upadhyay AK, Cheng X. Dynamics of histone lysine methylation: structures of methyl writers and erasers. Prog. Drug Res. 67, 107–124 (2011).
    • 95. Martin C, Zhang Y. The diverse functions of histone lysine methylation. Nat. Rev. Mol. Cell. Biol. 6(11), 838–849 (2005).
    • 96. Wang H, Guo R, Du Z et al. Epigenetic targeting of granulin in hepatoma cells by synthetic CRISPR dCas9 Epi-suppressors. Mol. Ther. Nucleic Acids 11, 23–33 (2018).
    • 97. Kim JM, Kim K, Schmidt T et al. Cooperation between SMYD3 and PC4 drives a distinct transcriptional program in cancer cells. Nucleic Acids Res 43(18), 8868–8883 (2015).
    • 98. Hombach S, Kretz M. Non-coding RNAs: classification, biology and functioning. Adv. Exp. Med. Biol. 937, 3–17 (2016).
    • 99. Jain S, Thakkar N, Chhatai J, Pal Bhadra M, Bhadra U. Long non-coding RNA: functional agent for disease traits. RNA Biol. 14(5), 522–535 (2017).
    • 100. Jarroux J, Morillon A, Pinskaya M. History, discovery, and classification of lncRNAs. Adv. Exp. Med. Biol. 1008, 1–46 (2017).
    • 101. Engreitz JM, Haines JE, Perez EM et al. Local regulation of gene expression by lncRNA promoters, transcription and splicing. Nature 539(7629), 452–455 (2016).
    • 102. Zhu S, Li W, Liu J et al. Genome-scale deletion screening of human long non-coding RNAs using a paired-guide RNA CRISPR-Cas9 library. Nat. Biotechnol. 34(12), 1279–1286 (2016).
    • 103. Joung J, Engreitz JM, Konermann S et al. Genome-scale activation screen identifies a lncRNA locus regulating a gene neighbourhood. Nature 548(7667), 343–346 (2017).
    • 104. Liu Y, Cao Z, Wang Y et al. Genome-wide screening for functional long noncoding RNAs in human cells by Cas9 targeting of splice sites. Nat. Biotechnol. doi: 10.1038/nbt.4283 (2018).
    • 105. Liu SJ, Horlbeck MA, Cho SW et al. CRISPRi-based genome-scale identification of functional long noncoding RNA loci in human cells. Science 355(6320), (2017).
    • 106. Chen W, Zhang G, Li J et al. CRISPRlnc: a manually curated database of validated sgRNAs for lncRNAs. Nucleic Acids Res. 47(D1), D63–D68 (2019).
    • 107. Morelli E, Gulla A, Amodio N et al. CRISPR interference (CRISPRi) and CRISPR Activation (CRISPRa) to explore the oncogenic lncRNA network. Methods Mol. Biol. 2348, 189–204 (2021).
    • 108. Arnan C, Ullrich S, Pulido-Quetglas C et al. Paired guide RNA CRISPR-Cas9 screening for protein-coding genes and lncRNAs involved in transdifferentiation of human B-cells to macrophages. BMC Genomics 23(1), 402 (2022).
    • 109. Zhang B, Ye Y, Ye W et al. Two HEPN domains dictate CRISPR RNA maturation and target cleavage in Cas13d. Nat. Commun. 10(1), 2544 (2019).
    • 110. Liu L, Li X, Ma J et al. The molecular architecture for RNA-guided RNA cleavage by Cas13a. Cell 170(4), 714–726.e710 (2017).
    • 111. Tan MH, Li Q, Shanmugam R et al. Dynamic landscape and regulation of RNA editing in mammals. Nature 550(7675), 249–254 (2017).
    • 112. Savva YA, Rieder LE, Reenan RA. The ADAR protein family. Genome Biol. 13(12), 252 (2012).
    • 113. Splinter E, de Laat W. The complex transcription regulatory landscape of our genome: control in three dimensions. EMBO J. 30(21), 4345–4355 (2011).
    • 114. Zhou HY, Katsman Y, Dhaliwal NK et al. A Sox2 distal enhancer cluster regulates embryonic stem cell differentiation potential. Genes Dev. 28(24), 2699–2711 (2014).
    • 115. Canver MC, Lessard S, Pinello L et al. Variant-aware saturating mutagenesis using multiple Cas9 nucleases identifies regulatory elements at trait-associated loci. Nat. Genet. 49(4), 625–634 (2017).
    • 116. Gasperini M, Hill AJ, McFaline-Figueroa JL et al. A genome-wide framework for mapping gene regulation via cellular genetic screens. Cell 176(1–2), 377–390.e319 (2019).
    • 117. Gasperini M, Findlay GM, McKenna A et al. CRISPR/Cas9-mediated scanning for regulatory elements required for HPRT1 expression via thousands of large, programmed genomic deletions. Am. J. Hum. Genet. 101(2), 192–205 (2017).
    • 118. Sanjana NE, Wright J, Zheng K et al. High-resolution interrogation of functional elements in the noncoding genome. Science 353(6307), 1545–1549 (2016). •• First paper that utilized CRISPR screens to identify regulatory elements involved in cancer drug resistance.
    • 119. Liu X, Zhang Y, Chen Y et al. In situ capture of chromatin interactions by biotinylated dCas9. Cell 170(5), 1028–1043.e1019 (2017).
    • 120. Rajagopal N, Srinivasan S, Kooshesh K et al. High-throughput mapping of regulatory DNA. Nat. Biotechnol. 34(2), 167–174 (2016).
    • 121. Fulco CP, Munschauer M, Anyoha R et al. Systematic mapping of functional enhancer-promoter connections with CRISPR interference. Science 354(6313), 769–773 (2016).
    • 122. Korkmaz G, Lopes R, Ugalde AP et al. Functional genetic screens for enhancer elements in the human genome using CRISPR-Cas9. Nat. Biotechnol. 34(2), 192–198 (2016).
    • 123. Ren X, Wang M, Li B et al. Parallel characterization of cis-regulatory elements for multiple genes using CRISPRpath. Sci. Adv. 7(38), eabi4360 (2021).
    • 124. Hartenian E, Doench JG. Genetic screens and functional genomics using CRISPR/Cas9 technology. FEBS J 282(8), 1383–1393 (2015).
    • 125. Klann TS, Black JB, Chellappan M et al. CRISPR-Cas9 epigenome editing enables high-throughput screening for functional regulatory elements in the human genome. Nat. Biotechnol. 35(6), 561–568 (2017). • Used dCas9-based screens to identify novel tumor type-specific regulatory elements using unique tools (CRISPRi and CRISPRa).
    • 126. Xie S, Duan J, Li B, Zhou P, Hon GC. Multiplexed engineering and analysis of combinatorial enhancer activity in single cells. Mol. Cell. 66(2), 285–299.e285 (2017).
    • 127. Fulco CP, Nasser J, Jones TR et al. Activity-by-contact model of enhancer-promoter regulation from thousands of CRISPR perturbations. Nat. Genet. 51(12), 1664–1669 (2019).
    • 128. Zheng H, Xie W. The role of 3D genome organization in development and cell differentiation. Nat. Rev. Mol. Cell. Biol. 20(9), 535–550 (2019).
    • 129. Vertii A. Stress as a chromatin landscape architect. Front. Cell. Dev. Biol. 9, 790138 (2021).
    • 130. Tang WY, Ho SM. Epigenetic reprogramming and imprinting in origins of disease. Rev. Endocr. Metab. Disord. 8(2), 173–182 (2007).
    • 131. Mehmood R, Varga G, Mohanty SQ et al. Epigenetic reprogramming in Mist1 (-/-) mice predicts the molecular response to cerulein-induced pancreatitis. PLOS ONE 9(1), e84182 (2014).
    • 132. Heintzman ND, Stuart RK, Hon G et al. Distinct and predictive chromatin signatures of transcriptional promoters and enhancers in the human genome. Nat. Genet. 39(3), 311–318 (2007).
    • 133. Visel A, Rubin EM, Pennacchio LA. Genomic views of distant-acting enhancers. Nature 461(7261), 199–205 (2009).
    • 134. Ong CT, Corces VG. Enhancer function: new insights into the regulation of tissue-specific gene expression. Nat. Rev. Genet. 12(4), 283–293 (2011).
    • 135. Eggermann T, Schonherr N, Meyer E et al. Epigenetic mutations in 11p15 in Silver-Russell syndrome are restricted to the telomeric imprinting domain. J. Med. Genet. 43(7), 615–616 (2006).
    • 136. Ballestar E, Esteller M. Epigenetic gene regulation in cancer. Adv. Genet. 61, 247–267 (2008).
    • 137. Ng JM, Yu J. Promoter hypermethylation of tumour suppressor genes as potential biomarkers in colorectal cancer. Int. J. Mol. Sci. 16(2), 2472–2496 (2015).
    • 138. Wu J, He K, Zhang Y et al. Inactivation of SMARCA2 by promoter hypermethylation drives lung cancer development. Gene 687, 193–199 (2019).
    • 139. Wang Q, Dai L, Wang Y et al. Targeted demethylation of the SARI promotor impairs colon tumour growth. Cancer Lett. 448, 132–143 (2019).
    • 140. Saunderson EA, Stepper P, Gomm JJ et al. Hit-and-run epigenetic editing prevents senescence entry in primary breast cells from healthy donors. Nat. Commun. 8(1), 1450 (2017).
    • 141. Kardooni H, Gonzalez-Gualda E, Stylianakis E, Saffaran S, Waxman J, Kypta RM. CRISPR-mediated reactivation of DKK3 expression attenuates TGF-beta signaling in prostate cancer. Cancers (Basel) 10(6), (2018).
    • 142. Yoshida M, Yokota E, Sakuma T et al. Development of an integrated CRISPRi targeting DeltaNp63 for treatment of squamous cell carcinoma. Oncotarget 9(49), 29220–29232 (2018).
    • 143. Hu Y, Zhang H, Guo Z et al. CKM and TERT dual promoters drive CRISPR-dCas9 to specifically inhibit the malignant behavior of osteosarcoma cells. Cell Mol. Biol. Lett. 28(1), 52 (2023).
    • 144. Xu X, Li J, Zhu Y et al. CRISPR-ON-mediated KLF4 overexpression inhibits the proliferation, migration and invasion of urothelial bladder cancer in vitro and in vivo. Oncotarget 8(60), 102078–102087 (2017).
    • 145. Pakalniskyte D, Schonberger T, Strobel B et al. Rosa26-LSL-dCas9-VPR: a versatile mouse model for tissue specific and simultaneous activation of multiple genes for drug discovery. Sci. Rep. 12(1), 19268 (2022).
    • 146. Garcia-Bloj B, Moses C, Sgro A et al. Waking up dormant tumor suppressor genes with zinc fingers, TALEs and the CRISPR/dCas9 system. Oncotarget 7(37), 60535–60554 (2016).
    • 147. Escriva-Fernandez J, Cueto-Urena C, Solana-Orts A, Lledo E, Ballester-Lurbe B, Poch E. A CRISPR interference strategy for gene expression silencing in multiple myeloma cell lines. J. Biol. Eng. 17(1), 34 (2023).
    • 148. Okano H, Morimoto S. iPSC-based disease modeling and drug discovery in cardinal neurodegenerative disorders. Cell Stem Cell 29(2), 189–208 (2022).
    • 149. Li Y, Li L, Chen ZN, Gao G, Yao R, Sun W. Engineering-derived approaches for iPSC preparation, expansion, differentiation and applications. Biofabrication 9(3), 032001 (2017).
    • 150. Mandai M, Watanabe A, Kurimoto Y et al. Autologous induced stem-cell-derived retinal cells for macular degeneration. N. Engl. J. Med. 376(11), 1038–1046 (2017).
    • 151. Michurina S, Stafeev I, Boldyreva M et al. Transplantation of adipose-tissue-engineered constructs with CRISPR-mediated UCP1 activation. Int. J. Mol. Sci. 24(4), (2023).
    • 152. Lundh M, Plucinska K, Isidor MS, Petersen PSS, Emanuelli B. Bidirectional manipulation of gene expression in adipocytes using CRISPRa and siRNA. Mol. Metab. 6(10), 1313–1320 (2017).
    • 153. Giehrl-Schwab J, Giesert F, Rauser B et al. Parkinson's disease motor symptoms rescue by CRISPRa-reprogramming astrocytes into GABAergic neurons. EMBO Mol. Med. 14(5), e14797 (2022).
    • 154. Petazzi P, Torres-Ruiz R, Fidanza A et al. Robustness of catalytically dead Cas9 activators in human pluripotent and mesenchymal stem cells. Mol. Ther. Nucleic Acids 20, 196–204 (2020).
    • 155. Horike S, Mitsuya K, Meguro M et al. Targeted disruption of the human LIT1 locus defines a putative imprinting control element playing an essential role in Beckwith–Wiedemann syndrome. Hum. Mol. Genet. 9(14), 2075–2083 (2000).
    • 156. Camprubi C, Coll MD, Villatoro S et al. Imprinting center analysis in Prader–Willi and Angelman syndrome patients with typical and atypical phenotypes. Eur. J. Med. Genet. 50(1), 11–20 (2007).
    • 157. Park JW, Han JW. Targeting epigenetics for cancer therapy. Arch. Pharm. Res. 42(2), 159–170 (2019).
    • 158. Vo BT, Kwon JA, Li C et al. Mouse medulloblastoma driven by CRISPR activation of cellular Myc. Sci. Rep. 8(1), 8733 (2018).
    • 159. Takahashi K, Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126(4), 663–676 (2006). •• This seminal paper identified a cocktail of factors that, when overexpressed in differentiated cells, can reprogram them to pluripotent state.
    • 160. Liu XS, Wu H, Krzisch M et al. Rescue of fragile X syndrome neurons by DNA methylation editing of the FMR1 gene. Cell 172(5), 979–992.e976 (2018).
    • 161. Bengtsson NE, Hall JK, Odom GL et al. Corrigendum: muscle-specific CRISPR/Cas9 dystrophin gene editing ameliorates pathophysiology in a mouse model for Duchenne muscular dystrophy. Nat. Commun. 8, 16007 (2017).
    • 162. Black JB, Adler AF, Wang HG et al. Targeted epigenetic remodeling of endogenous loci by CRISPR/Cas9-based transcriptional activators directly converts fibroblasts to neuronal cells. Cell Stem Cell 19(3), 406–414 (2016).
    • 163. Rots MG, Jeltsch A. Editing the epigenome: overview, open questions, and directions of future development. Methods Mol. Biol. 1767, 3–18 (2018).
    • 164. Policarpi C, Dabin J, Hackett JA. Epigenetic editing: dissecting chromatin function in context. Bioessays 43(5), e2000316 (2021).
    • 165. Wilson RC, Gilbert LA. The promise and challenge of in vivo delivery for genome therapeutics. ACS Chem. Biol. 13(2), 376–382 (2018).
    • 166. Gemberling MP, Siklenka K, Rodriguez E et al. Transgenic mice for in vivo epigenome editing with CRISPR-based systems. Nat. Methods 18(8), 965–974 (2021).
    • 167. Richards DY, Winn SR, Dudley S et al. AAV-mediated CRISPR/Cas9 gene editing in murine phenylketonuria. Mol. Ther. Methods Clin. Dev. 17, 234–245 (2020).
    • 168. Gao J, Bergmann T, Zhang W, Schiwon M, Ehrke-Schulz E, Ehrhardt A. Viral vector-based delivery of CRISPR/Cas9 and donor DNA for homology-directed repair in an in vitro model for canine hemophilia B. Mol. Ther. Nucleic Acids 14, 364–376 (2019).
    • 169. Staahl BT, Benekareddy M, Coulon-Bainier C et al. Efficient genome editing in the mouse brain by local delivery of engineered Cas9 ribonucleoprotein complexes. Nat. Biotechnol. 35(5), 431–434 (2017).
    • 170. Lee K, Conboy M, Park HM et al. Nanoparticle delivery of Cas9 ribonucleoprotein and donor DNA in vivo induces homology-directed DNA repair. Nat. Biomed. Eng. 1, 889–901 (2017).
    • 171. Qiu M, Glass Z, Chen J et al. Lipid nanoparticle-mediated codelivery of Cas9 mRNA and single-guide RNA achieves liver-specific in vivo genome editing of Angptl3. Proc. Natl Acad. Sci. USA 118(10), (2021).
    • 172. Mirjalili Mohanna SZ, Hickmott JW, Lam SL et al. Germline CRISPR/Cas9-mediated gene editing prevents vision loss in a novel mouse model of aniridia. Mol. Ther. Methods Clin. Dev. 17, 478–490 (2020).
    • 173. Saito M, Xu P, Faure G et al. Fanzor is a eukaryotic programmable RNA-guided endonuclease. Nature 620(7974), 660–668 (2023).