Next Article in Journal
Post-Bariatric Hypoglycemia Is Associated with Endothelial Dysfunction and Increased Oxidative Stress
Next Article in Special Issue
Therapeutic miR-506-3p Replacement in Pancreatic Carcinoma Leads to Multiple Effects including Autophagy, Apoptosis, Senescence, and Mitochondrial Alterations In Vitro and In Vivo
Previous Article in Journal
Phosphatidylethanolamine Deficiency and Triglyceride Overload in Perilesional Cortex Contribute to Non-Goal-Directed Hyperactivity after Traumatic Brain Injury in Mice
Previous Article in Special Issue
Impact of miR-1/miR-133 Clustered miRNAs: PFN2 Facilitates Malignant Phenotypes in Head and Neck Squamous Cell Carcinoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Elucidating miRNA Function in Cancer Biology via the Molecular Genetics’ Toolbox

1
Cluster of Regenerative Medicine, Advanced Medical and Dental Institute, Universiti Sains Malaysia, Bertam, Kepala Batas 13200, Pulau Pinang, Malaysia
2
Northern Institute for Cancer Research, Paul O’Gorman Building, Medical School, Framlington Place, Newcastle upon Tyne NE2 4HH, UK
3
Prinses Máxima Centrum Voor Kinderoncologie Heidelberglaan 25, 3584 CS Utrecht, The Netherlands
*
Author to whom correspondence should be addressed.
Submission received: 25 October 2021 / Revised: 23 November 2021 / Accepted: 30 November 2021 / Published: 15 April 2022
(This article belongs to the Special Issue MicroRNA in Solid Tumor and Hematological Diseases 2.0)

Abstract

:
Micro-RNA (miRNAs) are short non-coding RNAs of about 18–20 nucleotides in length and are implicated in many cellular processes including proliferation, development, differentiation, apoptosis and cell signaling. Furthermore, it is well known that miRNA expression is frequently dysregulated in many cancers. Therefore, this review will highlight the various mechanisms by which microRNAs are dysregulated in cancer. Further highlights include the abundance of molecular genetics tools that are currently available to study miRNA function as well as their advantages and disadvantages with a special focus on various CRISPR/Cas systems This review provides general workflows and some practical considerations when studying miRNA function thus enabling researchers to make informed decisions in regards to the appropriate molecular genetics tool to be utilized for their experiments.

1. MicroRNA: A Brief Introduction

MicroRNAs were first discovered almost three decades ago by Lee et al. (1993) when his research group discovered that the lin-4 transcripts of 22 and 61 nucleotide (nt) bases long, known at that time as short RNA, controlled the expression of a lin-14 protein which is crucial for embryonic development of C. elegans [1]. Sequence study of the lin-4 gene revealed that repeat elements located within the 3′-untranslated region (3′-UTR) are complementary to sequences found inside both of the lin-4 transcripts. This group hypothesized that an RNA–RNA hybrid formation may contribute to the downregulation of lin-14 mRNA, although the complementary sequence stretches only 10 nt base long. Through this landmark research, the term microRNA was coined to describe short RNA oligos of approximately 18 to 20 nt long which functions to regulate protein expression at the transcriptome level [2,3,4].
Many cell types depend on the complex system of microRNAs to regulate their normal biological and physiological functions [5]. However, when these networks are dysregulated, normal cells could acquire oncogenic traits which drive conversion into malignant counterparts [6,7]. Among the common mechanisms which facilitate tumorigenicity include amplification of loci, various classes of mutations, epigenetic modifications on microRNA loci and alterations to transcription factors controlling microRNA expression [8,9]. The accumulation of these changes progressively lead to carcinogenesis and over time bring rise to the various hallmarks of cancer which includes conferring selective proliferative advantage, altered stress response, increased vascularization, enhancement of invasion and metastasis, rewiring of metabolic pathways, evasion of immune response and augmentation of the tumor microenvironment [10,11].
In order to elucidate the various effects of microRNA dysregulation in cancers, a wide array of novel technologies in molecular genetics are required. The rapid progress of discoveries in molecular biology has led to innovative applications which are capable of deciphering mechanisms of disease onset and progression with great precision where of late, research related to microRNA regulation in cancers have incorporated various pioneering molecular techniques including miRNA sponges, antisense oligonucleotides (antagomirs) and miRNA mimics [12,13,14].
Moreover, the advent of computational biology has led to the development of various databases and tools allowing for miRNA target analyses [15]. Target identification is crucial for the translation of miRNA potential activity as protein function predicts the implication of miRNA dysregulation in disorders [16,17]. Tools such as miRDB, TargetScan, mirTarBase and mirWalk, which are openly available online with ease of access and user-friendly interfaces, would significantly aid researchers in identifying a probable target for studied miRNA [18,19,20,21].
More recently, CRISPR has emerged as a game-changing tool in this field bolstering a range of derivative Cas proteins which are capable of interrogating and augmenting different microRNA functions in cancers [22,23,24]. This review will focus to detail the use of various cutting-edge molecular genetics technologies in elucidating the microRNA dysregulation of cancers to provide a clearer perspective on the advantages and limitations of each technique as well as a comparison of their utility in individual studies and delineate practical considerations in the implementation of new strategies such as CRISPR [25,26].

2. Mechanisms of microRNA Dysregulation in Cancer

Similar to genes, microRNA are also subjected to dysregulation, therefore various mechanisms need to be considered when studying miRNA dysregulation. Amongst them are the amplification of loci which leads to the overexpression of amplified genes, mutation leading to deregulated or loss of function, epigenetic control contributing to gene activation or suppression and the effects of trans activating elements on gene expressional control.

2.1. Amplification of Loci

MicroRNA loci amplification is a common occurrence in various cancers (Figure 1). Amplification of the genomic region often causes overexpression of the miRNAs thus resulting in the depletion of its target mRNA [27]. This in turn leads to dysregulation in critical signaling pathways which contribute to cancer development and progression.
This can be observed in embryonic brain tumors where the C19 miRNA cluster was amplified [28,29]. Interestingly, the same research also showed that the C19 cluster was fused to the TTYH1 gene, which further drives the expression of the amplified C19 cluster polycistronically. This causes extreme overexpression of DNM3TB which is an effector of the C19 miRNA cluster target, RBL2. This study clearly shows that microRNA loci amplification leads to the dysregulation of target genes and their effectors.
Amplification also occurred in the miR-17-92 cluster where it was found that the miRNAs within the cluster were significantly overexpressed in human lung cancer [30]. MiR-17-92 cluster copy number was significantly higher when compared to the copy number of miR-106a-92 and miR 106b-25. Additionally, the same cluster was also found to be amplified in acute myeloid leukemia with MLL rearrangement where it was found that this cluster affects cellular differentiation and proliferation [31].
Furthermore, miR-23a locus was also found to be implicated in the same manner in gastric cancer. Research found that the miRNA was over-expressed and that the miRNA gene was amplified along with other miRNAs (miR-1274a, miR-196b, miR-4298, miR-181c, miR-181d, miR-23a, miR-27a and miR-24-2) at loci 19p13.13. The upregulation of these miRNAs contributed to gastric cancer progression. Additionally, the group also revealed that miR-23a downregulated MT2A and that inhibition of the microRNA halted tumor growth [32].
It was also reported that amplification of miR-191/425 promotes the growth of breast cancer [33]. The group reported that the copy number of miR-191/425 was significantly higher in primary breast tumor in comparison to normal tissue. MicroRNA-191/425 target DICER1 which leads to the global miRNA dysregulation due to its function in processing pre-miRNA into mature miRNA, thus effectively leading to the promotion of breast cancer tumorigenesis.

2.2. Mutation (Single Point Mutation, Deletion, Insertion, Base Substitution)

Mutation to genes such as substitution or deletion often leads to the production of nonfunctional translated proteins due to impaired codon arrangements [34,35,36]. As a consequence, critical cellular signaling pathways are frequently dysregulated due to the impairment in protein structure and function where this culminates with the development of various cancers [37,38,39,40]. A similar phenomenon can be observed in miRNA where a mutation in the miRNA locus would lead to dysfunctional miRNA expression (Figure 2).
This was proven in the case of the miR-15a and miR-16-1 locus where partial homozygous deletion of the locus led to the depletion of miR-15a expression. Furthermore, the group also reported a single base substitution of T to C 21bp upstream of the miR-15a coding sequence. It was postulated from this research that these mutations lead to the development of prostate cancer due to the prevalence of these mutations in xenografts [41].
Furthermore, point mutation of miRNA was also found to play a hand in leukemogenesis. It was found that point mutation in the miR-142 resulted in the loss of function of that miRNA. This potentiates AML towards myeloid lineages while suppressing lymphoid lineages differentiation. The loss of miR-142 leads to the accumulation of ASHL1 resulting in the maintenance of the HOXA genes. This caused the myeloid progenitors to be locked in the myeloid stage thus preventing proper cellular differentiation [42]. Additionally, it was also reported from cancer cohort (TCGA) data that the miR-142-3p seed region harbors the mutation at the 7th and 6th nucleotide where A > G and C > G/T occurs predominantly in diffuse large b-cell lymphoma [43]. Earlier results reported also that mutation in the miR-142-3p seed sequence shifted miR-142 targeting from RAC1 towards ZEB2 in diffuse large b-cell lymphoma which could attribute to disease progression [44].

2.3. Epigenetic Modifications on miRNA Loci

Epigenetic modulation plays a critical role in the regulation of miRNA expression [45,46]. Epigenetic markers such as DNA methylation, histone state, an abundance of CpG sites and binding of epigenetic modulators influence the accessibility of miRNA loci to various transcription machinery (Figure 3A,B). In this case, a group reported that PPARγ binds consensus sequence located on the miR-122 promoter to promote expression by inducing histone acetylation. They reported that induction of hepatocellular carcinoma cells with 5′aza-2′deoxycytidine (a compound that inhibits methyltransferase) prevented the N-cor/SMRT/HMeT complex, which is a miR-122 inhibitor from binding onto the consensus promoter sequence which in turn caused a reduction in H3K9 demethylation (heterochromatin mark on the genome, usually found to be abundant in repressed genes) and provided an accessible miR-122 locus which allowed PPARγ to bind predominantly [47].
Histone state plays a major role in the accessibility of crucial transcription machinery onto gene locus which triggers transcription or suppression. Depending on the state of chromatin brought about by the actions of histone modifiers genes can either be activated or inactivated. In miRNA regulation, it was reported that the interplay between miRNA and histone modifiers exists in modulating disease progression [48,49,50]. It was found that miR-125a-5p expression negatively regulates HDAC5 in breast cancer which contributes to disease progression via a feedback loop. The study found that miR-125a-5p binds the 3′-UTR of HDAC5 which leads to the reduction in HDAC5 dependent RUNX3 acetylation on p300/RUNX3 complex resulting in the dissociation of the complex binding on the miR-125a-5p locus. Using HDAC inhibitor, HDAC5 activity was suppressed leading to the reduction of miR-125a-5p which leads to apoptosis suggesting interplay between miRNA and HDACs in mediating apoptotic evasion [51]. In addition, it was also reported that the EZH2 histone modifier, which induces heterochromatinization, was overexpressed in pancreatic duct adenocarcinoma (PDAC) leading to the suppression of miR-218. Direct EZH2 siRNA targeting leads to restoration in miR-218 expression as this abrogates EZH2 binding onto the miRNA locus also blocking the recruitment of DNMTs via EZH2 [52].
Additionally, it was also shown that CpG sites, influence the expression of miR-335 located within the MEST gene locus downstream of MEST exon 1 which may contribute to hepatocarcinogenesis. In this study, it was found that the CpG-rich site located upstream of the miR-335 beyond exon 1 of MEST resulted in the repression of miR-335 expression due to DNA hypermethylation. The group also reveals that treatment with 5′aza-2′deoxycytidine, a known DNA demethylator, resulted in the upregulation of the suppressed miR-335 [53].
It was found that in prostate cancer overexpression of the MED1 gene would lead to the disease progression [54,55]. This effect could be mitigated via MED1 suppression by miR-205. However, it was found that the miR-205 locus was hypermethylated in prostate cancer thus leading to the epigenetic silencing of the miRNA. Additionally, the same group found that H3K9Ac, a histone mark for euchromatin (highly accessible areas within the genome usually susceptible to transcription factor binding), was significantly reduced on miR-205 loci. Furthermore, they also reported the prevalence of epigenetic in various cancer cells and found that miR-205 expression was inversely correlated with DNA methylation [56,57].
Results have shown that miR-195 was epigenetically influenced as well. It was found that the expression of miR-195, a tumor suppressor miRNA that is required for normal colon function, was downregulated in colorectal cancer cells [58,59]. It was also shown in the study that CpG islands upstream of the miR-195 locus were partially methylated in normal colorectal cells. However, in colorectal cancer, this site was fully methylated [60]. The findings from these studies suggest that DNA hypermethylation of miRNA loci in colorectal cancer plays a crucial role in oncogenesis by interfering with the tumor-suppressive activity of miR-195.
Reports have also shown that miR-340 is implicated epigenetically in neuroblastoma [61,62]. CpG-rich sites were found downstream of the predicted transcription start site (TSS) of miR-340 and these sites were extensively methylated. Interestingly, these sites were also found to overlap with RNA polymerase II (pol II) binding sites thus suggesting epigenetic modulation in transcriptional control of miRNA [63]. Treatment with 5′aza-2′deoxycytidine reduced methylation of the CpG sites which in turn led to an increase in miR-340 activity via reduction of its target gene SOX2. These findings clearly highlight the role of epigenetic modifications in miRNA dysregulation.

2.4. Transcription Factors (TF) Controlling miRNA Expression Dysregulation

Transcription factors (TFs) are well known to either promote or suppress gene expression [64]. Transcription factors are generally understood as proteins that bind specific consensus sequences located on genomic regions, particularly, on gene loci [65,66]. The binding of TFs can either occur proximal or distal from the TSS [67,68]. TFs that promote accessibility on the genomic region for RNA pol II binding and other TFs are known as activators whereas suppressors are TFs that promote inaccessibility [69]. The mechanism of action of TFs involves the recruitment of other TFs or epigenetic modulators on specific genomic loci. Depending on what type of TF bounded (Suppressors or Activators) they partner with different epigenome modulators (HDAC, HMET, HAT, p300). This leads to either upregulation or downregulation of genes affected [70,71].
Proper TF function is crucial for the regulation of miRNA expression. This can clearly be seen in the interaction between the transcription factor ETS5 and miR-155. It was shown that ETS5 binds to miR-155 TSS and activates the expression of the pri-miR-155 which in turn leads to lipopolysaccharide activation in the human embryonic kidney. Contrarily, the expression of miR-155 was downregulated by the addition of IL-10, an inhibitor of ETS5. This is an example of an activator mechanism in miRNA expression [72].
Additionally, it was also found that GATA3 modulates the activity of miR-29b by binding on the promoter of miR-29b [73]. Reporter assay also showed induction in expression when miR-29b loci containing the GATA3 binding sequences were cloned into luciferase plasmids [74]. However, deletion of the promoter binding sequence resulted in the reduction in luciferase activity clearly showing that GATA3 binding is crucial for miR-29b induction.
Interestingly, it was also reported that miR-1 and miR-206 were both regulated by NRF2 transcription factors in an HDAC4-dependent manner [75,76]. In this study, downregulation of NRF2 resulted in an increase in miR-1 and miR-206 expression. This indicates that the TF acts as a suppressor. NRF induces HDAC activity resulting in a reduction in accessibility (heterochromatin) on the genomic loci, thus inactivating transcription. Furthermore, downregulation of HDAC resulted in an increase in miR-1 and miR-206 expression.
A summary of the various mechanisms involved in miRNA dysregulation is presented in Table 1.

3. Conventional Molecular Genetics Methods for Studying miRNA Dysfunction

Currently, a myriad of molecular genetic techniques is available to study the various phenomenon that cause miRNA dysregulation. (Figure 4). The subsequent portions of this review will describe each method in detail.

3.1. MicroRNA Sponges

In essence, miR-Sponges sequester miRNAs by overexpressing their target complementary 3′-UTR sequences using a plasmid-based system where the 3′-UTR is expressed using a strong promoter, e.g., the CMV promoter, basically, these plasmids overexpressed the miRNA target sequence in which miRNA binds thus “pull” the miRNAs away from their target sequence [87,88]. The expression of the miRNA target sequences act by providing an accessible template for the target miRNA to bind, this, in turn, will block the miRNA from binding to target messenger RNA as the miRNA has a higher affinity for overexpressed, readily available, target site. This leads to the knockdown effect of targeted miRNAs as sponges sequester miRNA from binding to its actual target, resulting in the accumulation of the mRNA. This system usually fuses the 3′-UTR (miRNA target sequence) at the 3′ end of a reporter gene such as GFP or luciferase [89,90]. A CMV promoter plasmid drives the reporter gene expression bearing the miRNA target sequence on its 3′-UTR constitutively. The sequestering of the targeted miRNAs can be seen via a reduction in reporter gene signal in comparison to control which expresses reporters without the miRNA target sequence.
Studies also showed that endogenous sponges or “natural sponges” act in the same manner where they regulate miRNA function. In this case, it was reported that long intergenic noncoding RNA, named linc-RNA-ROR, regulates miR-145 which is crucial in regulating the self-renewal capability of stem cells as miR-145 directly targets the mRNA of TFs crucial for self-renewal such as the SOX2, Oct-4 and Nanog. As mentioned previously, the group also employed the use of reporter gene bounded to miR-145 target sequences. Luciferase gene fused with linc-RNA-ROR, Oct-4, SOX2 and Nanog transcripts, respectively, were transfected along with a miR-145 mimic (double-stranded RNA molecules that mimic miR-145 function). This resulted in the reduction in the reporter gene, Oct-4, SOX2 and Nanog activity, thus further confirming that miR-145 binds 3′-UTR of these genes [91,92].
MicroRNA sponges are more accurate in terms of defining specific miRNA targets since the target is cloned directly into a reporter gene mRNA. This enables the ectopic expression of plasmid sponges and miRNA mimic, which allows for more accurate results in miRNA sequence target confirmation in specific samples. On the other hand, expression of the target miR-sponges natively may result in the false reduction of reporter signal as endogenous miRNA may contribute to reporter silencing by competing for binding on the overexpressed target sequence. Additionally, the use of lentiviral plasmid allows for a more stable integration of the reporter system in cells which would result in a stable expression of the miR-targets [93,94].
However, there are cases of off-target effects brought about when using miR-sponges, this is because some miRNAs have similar seed sequences [95]. These seed sequences allow multiple miRNAs to bind to a similar target sequence. This could further complicate matters as targeted knockdown of a single miRNA may lead to the effect of targeting multiple miRNAs sharing similar seed sequences. Furthermore, miRNA sponges would compete with endogenous mRNAs with complementary 3′-UTR towards the target miRNAs this would result in steric hindrance, thus requiring more miR-sponges to suppress endogenous miRNAs. Although this might seem probable with titrations, too much sponge in itself may result in off-target effects [96,97].

3.2. Antisense Oligonucleotides/Antagomirs

Antisense oligonucleotides or generally known as antagomirs are a short sequence of RNA usually 13–25nt long [98]. Antagomirs are similar to siRNA tools as they bind target RNA to block their function [99,100,101]. These RNAs are complementary and can form base pairing with their target miRNA [102,103]. This would lead to the blockade of the miRNA function where miRNA in the miRISC complex cannot form complementary base pair with its target mRNA at the 3′-UTR. ASOs are often chemically modified by adding the 2′-O-methyl group to increase their potency [104]. Additionally, a pH-sensitive peptide carrier (pHLIP peptide) was also fused to ASO, this improves cell entry via the non-endocytic route [105,106].
miRNA control by ASO can be seen in research by Sun and colleagues, 2015, where the group silenced miR-320 [107]. The silencing of miR-320 led to the proliferation of glioma cells suggesting that miR-320 expression is crucial for the anti-proliferative effect. Additionally, the group also discovered that the E2F1 transcription factor, a target of miR-320, was upregulated.
ASOs were also employed in the silencing of miR-30a-5p in gliomas. In this study, it was found that the miRNA controls SEPT7crucial for proper neurogenesis and that overexpression of miR-30a-5p reduces the abundance of SEPT7 leading to glioma development [108]. Additionally, ASOs was used in the treatment of leukemia where it targeted miR-126 which was found to be hyper expressed in patients with worse outcome. Targeting miR-126 resulted in the reduction of leukemic stem cells suggesting its critical role in leukemogenesis [109,110].
An advantage of using ASOs is that they are relatively easy to use; conceptually the ASOs will complementarily bind to its target sequence thus blocking the miRNA in the miR induced silencing complex (miRISC) from binding to its target mRNA. However, as simple as it may seem, the inhibition of target miRNAs requires the ASOs to be extensively modified [111]. Antagomirs can either be directly complementary or modified by adding modified ribonucleotides such as locked nucleic acid (LNA) a type of ribonucleotide with chemical bonds between the 2′oxygen and 4-carbon via a methylene bridge [112,113], ribonucleotides with phosphorothioate bonds [114] rather than phosphodiester bonds or nucleotide with modified 2′ position which can either be 2-O-methyl, 2-Omethoxyethyl or 2′-fluoro [115]. These modifications are crucial for the functioning of the antagomirs as the modifications would confer nuclease resistance and an increase in ASO potency. Generation of these modified ASOs involves complex chemical reactions despite being conceptually easy.
Furthermore, antagomirs usage interferes with downstream target mRNA quantification. Using qPCR-specific oligomers (primers or probes) designed against the target miRNAs, detection is based on the annealing of these oligomers onto the miRNA strand. Due to the formation of the ASO/miRNA heteroduplexes, these oligomers could compete for binding to the target miRNA resulting in a false positive reduction of miRNAs [116].

3.3. miRNA Mimic

Mimics, as the name suggests, are double-stranded or single-stranded RNAs that have a similar sequence to miRNAs (Figure 5). These types of RNA act as an overexpression tool where they mimic the function of respective miRNAs [117]. Relatively easy to deliver and use, this technique becomes widespread in studying the effects of miRNA overabundance and miRNA replacement [118].
An example was recent research using mimics for miR-15a and miR16-1 in chronic lymphoid leukemia [119]. Results obtained showed that the two miRNAs affect CLL viability and that overexpression of these miRNAs using mimic significantly reduced cell viability suggesting that the use of these mimics could improve leukemic outcomes. Additionally, mimic for miR-494 targeting survivin, a crucial TF for the expression of kit in gastrointestinal stromal tumors, was also used [120]. Using 3′-UTR of survivin bounded to luciferase reporter gene the group co-transfected HeLa cells with miR-494 mimic and found that luciferase activity relative to negative control mimic was suppressed. Luciferase activity was rescued using inhibitors toward the miR-494 mimic.
A miRNA mimic was also used to study the effects of miR-126-5p/3p in AML as it was found to be highly expressed in patients with lower overall survival. Apoptosis inhibition and hyperproliferation were observed following mimic treatment suggesting miR-126-3p/5p is an oncomiR. This microRNA serves as a prognostic marker where overexpression represents a poor prognosis. This study also found that miR-126-3p inhibits cytarabine inducing apoptosis in a patient following chemo [121]. Here we can see the use of mimic in inducing specific miRNA expression without the use of lengthy and complex transduction or plasmid-based systems.
However, it was found that the use of this mimic would lead to passenger strand loading onto the miRISC complex leading to the unwanted effect of the passenger strands towards complementary 3′-UTR of genes [122]. This would lead to false-positive changes and the phenotypic changes may not be attributed to the miR-mimic overexpression. Another study also reported that post-mir-mimic transfection, cells tend to accumulate RNA at high levels [123]. This would lead to undesirable effects as it was found that the accumulation of high molecular weight RNA led to disease progression. Furthermore, the same study also reported that treatment of cells at a high concentration of miRNA-mimic would lead to non-specific changes in gene expression patterns. Surprisingly, they also found that mimic transfection leads to the accumulation of mutated endogenous miRNA and that this may result in non-specific gene modulation.
A summary of the various conventional techniques currently available to study miRNA function is depicted in Table 2.

4. CRISPR/Clustered Regularly Interspaced Short Palindromic Repeats as an Emerging Molecular Genetic Tool to Study miRNA Dysregulation

The previous sections of this review have shown that studying miRNA dysregulation is crucial in understanding cancer biogenesis and development. It is also apparent that an array of molecular techniques is readily available currently to elucidate miRNA function. The advantages and common pitfalls associated with each technique were described extensively as well. Additionally, the advent of genome editing has presented us with, yet another technique known as Clustered Regularly Interspaced Short Palindromic Repeat (CRISPR) to study miRNA dysregulation. The utilization of this versatile technique to elucidate miRNA function will be the focus of this review henceforth.
In essence, CRISPR or Clustered Regularly Interspaced Short Palindromic Repeats is a defense mechanism of bacteria from invading DNA-phage [132,133]. Briefly, it is a mechanism of acquired immunity in bacteria where the DNA of invading phages will be incorporated between the repeating spacer units that will be transcribed into the CRISPR RNA (crRNA) whereupon secondary phage invasion, this DNA will act as a guide in cleaving the invading phage via Cas9 endonuclease [134]. The guide RNAs are complementary to the viral DNA sequence, thus the directing of Cas9 on target viral DNA is based on RNA-DNA direct complementarity.

4.1. CRISPR/Cas9 (CRISPR Associated Protein 9)

CRISPR/Cas9 is a revolutionary tool used initially by Feng Zhang’s group to mediate genome engineering [135]. This technique employs the use of Cas9 endonuclease and small guide RNA (sgRNA) as tools for targeted genome editing. The sgRNA would form a complex with the Cas9 endonuclease. A twenty-nucleotide sequence within the sgRNA adjacent to the protospacer adjacent motif (PAM) sequence at the 5′ end of the sgRNA construct would then act as the homing beacon in defining specific target sequence via complementary base pairing.
The complementary binding defined by the sgRNA sequence would then bring the Cas9 towards the target genomic site and the endonuclease would subsequently cleave the double-stranded DNA and form a double-strand break (DSB). This would then initiate a non-homologous end joining (NHEJ) process where DNA repair machinery would start to reattach the breaks. This repair mechanism results in mutations such as insertion, deletion and random base substitution due to the fact that NHEJ is error-prone. The resulting mutation would render the target gene dysfunctional. The endonuclease capability of Cas9 also enables double nicking. This was observed in the Cas9 mutant termed Cas9-D10A, single-stranded DNA cleavage was observed within proximity of the PAM site. Using Cas9-D10A dual sgRNA system targetting the plus and the minus strand could be co-transfected with Cas9-D10A allowing for single-stranded nicking producing overhangs [136]. Reattachment of the overhangs would introduce indels via NHEJ.
Besides DSB, the foreign sequence could also be added into the genome via homology-directed repair (HDR) [137]. HDR is a mechanism where DNA repair machinery would recognize a foreign DNA strand that has the same sequence homology with the broken strand and integrate that foreign strand into the breaks (DSBs). Specific DNA sequence can be placed inside the strand in between 5′ end and 3′end homology sequence where it can be incorporated into the repaired strand through the pairing of the introduced strand to the broken strand via complementary base pair of the homology sequence where subsequently the foreign strand will be ligated into the broken strand.
Utilizing the principle mentioned above (CRISPR/Cas9), one could either increase miRNA expression (HDR), do a knock-in study (HDR) or knockout the miRNA gene (NHEJ) directly on the genome level as this technique allows for highly specific genome editing. However, there are consequences of off target effects where the endonuclease would cleave non-targeted genomic loci. Various other CRISPR methods are available to mitigate this bottleneck and will be described subsequently.

4.1.1. CRISPR/Cas9 Direct Modulation of miRNA Genome

Several studies have shown the use of CRISPR in directly modulating miRNA genes [138,139]. This was shown to be effective in downregulating the expression of miR-17 [140]. Xi and group showed that designing small guide RNA targeting the 5′p and 3′p region of the pre-miRNA containing the dicer and drosha processing sites resulted in a significant reduction in the levels of mature miRNA expression and upregulation of the primary miRNA transcript. Additionally, it was also found that E2F1, a target of miR-17, was upregulated. PCR and subsequent sequencing of cleaved genomic target reveal different mutations which included deletions and insertions that were brought about by NHEJ on the miRNA loci.
Additionally, CRISPR/Cas9 was also used in depleting miR-210-3p in renal cell carcinoma [141]. Depletion targets the 3′p seed sequence and the stem-loop sequence of pre-miRNA-210-3p where Cas9 endonuclease cleavage resulted in the reduction of the mature miR-210-3p further showing that the CRISPR/Cas9 system can be used to directly target the genomic loci in which the miRNA gene is present.
However, certain complications could arise from using this technique since Cas9 cleavage results in indels due to the double-strand breaks generated. A common pitfall in using this technique is that if the PAM sequences are located inside the miRNA seed sequence, Cas9 cleavage could result in a mutant miRNA due to indels. Processing of the primary transcript would result in mutant miRNAs that target different genes and cause changes in phenotype, producing false-positive results. This is highly plausible as it was reported that a difference of four nucleotide bases could potentially cause off-target binding of sgRNA targeting for different miRNAs [140].

4.1.2. Cas9 Modulation: miRNA Gene Expression Control

Cas9 endonuclease cleavage of genomic loci is defined by the sgRNA. As such, the docking of Cas9 on a specific genome locus is based on sequence complementarity of the sgRNA and DNA forming the sgRNA-gDNA hybrid. This has allowed scientists to direct certain functional protein domains towards targeted gene loci via fusion with Cas9 to induce specific changes in gene expression [142,143,144]. This proof of concept was initially reported by Gersbach’s group in which they mutated the D10A mutant of Cas9 via site-directed mutagenesis and fused the “dead” Cas9 termed dCas9 to VP-64 a functional activator of gene expression [145]. The group also designed various enhancements on the dCas9 allowing for epigenetic modulation, gene repression, cell line-specific reprogramming, epigenetic reprogramming towards specific cells lines and endogenous gene expression control [146,147,148,149].
Using the dCas9-VP-64, Gersbach and colleagues designed sgRNAs targeting the promoter of IL1RN. Plasmid cassettes containing the dCas9 and sgRNA were then transfected into HEK293T. RNAseq and sub sequential DESEq analyses revealed only upregulation of IL1RN thus suggesting that the technique is highly specific. The research was also carried out by targeting the promoters of NANOG, MYOD, TERT, VEGFA, IL1B and HBG1 and 2. The results show significant upregulations of these genes. They also reported that binding of the dCas9-activator complex to enhancers and distal promoters could also activate gene expression [145].
Owing to the fact that crRNA-guided gene activation is specific as dCas9 could carry specific transactivation domains, Gersbach and colleagues also fused dCas9 with deactivators [150]. In this case, KRAB (Krüppel-associated box) domain was utilized to target a distal enhancer of the HBG gene. Results showed a significant reduction in HBG1 and HBG2 expression and an increase in heterochromatin signature indicating repression in the enhancer site. This interferes with the binding of native transcription factor onto the targeted loci bounded by dCas9-KRAB, leading to gene repression.
Using the same concept; theoretically, we can apply these techniques in inducing or repressing miRNA. This can be carried out by designing the sgRNA to be specific against a miRNA locus. Using the dCas9-VP64 activation tool, Gersbach induces expression of the targeted gene via multiple sgRNA transfections. This allows multiple binding of the fusion product to a proximal sequence within the promoter which will increase the VP-64 docking via dCas9 onto gene promoters resulting in a significant increase in gene expression. This could be employed by using plasmid cassettes which allow for the expression of multiple sgRNAs. Ventura’s lab utilized pX33 (addgene) specifically for in vivo application this plasmid enabled multiple sgRNA expression. This significantly increased VP-64 docking onto the miRNA gene locus thus activating them [151]. On the other hand, pX330A-1X6 (Addgene) from Yamamoto’s lab can be employed for in vitro applications. This plasmid allows for the expression of up to seven different sgRNAs which enables multiple dCas9-VP64 docking onto target sites which would increase miRNA expression. This is especially useful for a gain of function study [152]. However, for efficient gene induction, VP-64 must be bounded onto the gene locus at higher copy numbers. This could increase the chances of off-target due to the introduction of multiple sgRNAs.
Additionally, miRNA expression could also be repressed via dCas9-KRAB and sgRNA that targets specific miRNA locus. In this case, the KRAB domain induces epigenetic changes leading to heterochromatin by inducing histone trimethylation. This will result in a “tight” coil formation of DNA surrounding histones which would impair the accessibility of crucial transcription machinery onto their respective binding sites thus suppressing gene expression [153]. Owing to this fact, KRAB could be directed to specific miRNA genomic locus where the formation of heterochromatin would impair crucial transcription machinery docking onto miRNA locus thus suppressing miRNA expression [154,155]. The plasmid cassette that can for utilized for this purpose is the pLV hU6-sgRNA hUbC-dCas9-KRAB-T2a-Puro (Addgene) from the Gersbach lab. This plasmid could be used to co-express sgRNA along with dCas9-KRAB [150].
Proof of concept in the utilization of dCas9 for miRNA targeting was carried out by Zhao and colleagues where they designed sgRNA targeting the 5′ upstream sequence of miR-17-92 cluster loci [156]. This results in the reduction of mature miR19a, miR20 and miR-92 expression suggesting that dCas9 could be used to block the binding of crucial TFs thus blocking transcription elongation of pre-miRNA transcript. Therefore, using dCas9 alone without KRAB may suppress miRNA expression. However, it is highly probable that epigenetic modulation brought about by the KRAB domain would improve the efficiency of knockdown [157].

4.2. CRISPR/Cpf1 (CRISPR Associated Endonuclease in Prevotella and Francisella 1)

CRISPR/Cpf1 is a type II system that targets specific cleavage via RNA-guided docking of Cpf1 instead of Cas9 [158]. The Cpf1 system is more straightforward as compared to the Cas9 system in principle. Cpf1 1 requires a crRNA of approximately 42 nucleotides in length only and no complexing with tracrRNA is required. Furthermore, it was found that plasmid constructs containing only Cpf1 and sgRNA sequences are enough to induce DNA cleavage, thus making this system relatively easier than Cas9. However, the cleavage site of Cpf1 is located further away from the PAM site.
The advantage of using this system as compared to the Cas9 system is the generation of sticky ends, which would facilitate NHEJ mediated gene insertion. Since Cpf1 endonuclease activity resulted in 5bp overhangs, double-stranded DNA sequence carrying desired genes with flanking complementary sites towards the 5bp overhangs would improve gene knock-in via base pairing between the 5bp overhangs and gene inserts [159,160]. However, the suppression of microRNA demands more technical procedures as Cpf1 delivery would require both the endonuclease and the sgRNA to be delivered simultaneously. This can be observed in the lentiviral delivery of the Cpf1 system where plasmids carrying the endonuclease and the sgRNA would require genomic integration for stable gene knockout as in Cas9. Therefore, this approach could be technically more demanding than the delivery of antagomirs. This system also comes with an activator, however, it is currently used to target endogenous genes [161,162]. No direct targeting of miRNA by Cpf1 was reported to date, however, with the availability of Cpf1 tools for activation, we could induce acetylation of histones on miRNA locus activating miRNA expression as miRNA promoter was found to be associated with high histone acetylation state [163]. Thus far, using Cpf1 against miRNA is still lacking due to Cas9 still being the golden rule. However, the use of Cpf1 to suppress miRNA could be a potential field for applications development.

4.3. Cas12b/C2C1

Similar to Cpf1, Cas12b contains a single nuclease domain the RuvC domain making it smaller than Cas9. However, Cas12b requires tracrRNA to be incorporated for genomic cleavage this mechanism is similar to that of Cas9. PAM site for Cas12b is TTN compared to Cpf1 TTT. Engineering of the C2C1 by substituting amino acid residues at three different sites on the RuvC domain confers higher genomic cleavage efficiency and higher specificity as the percentage of perfect edits was found to be higher than that of Cas9 [164].
This Cas subtype may be more suitable in miRNA targeting as it has higher edit specificity. This should reduce the formation of off-target miRNAs resulting from Cas genomic cleavage.

4.4. CRISPR/Cas13

Cas13 or originally known as C2C2 was found to mediate direct RNA cleavage. There are various subfamilies of Cas13 consisting of Cas13a, Cas13b, Cas13c and Cas13d [165]. Cas13 endoribonuclease activity does not require PAM sequences and it was reported that direct complementarity towards target sequences via RNA-RNA hybrid is sufficient [166,167]. Furthermore, a small guide RNA sequence of 28nucleotide with flanking 36 nucleotide spacer was found to be sufficient for Cas13 mediated cleavage of mRNA where efficiency was comparable to that of shRNA mediated targeting with much higher specificity as shown by RNAseq data [167]. Although efficient, this technique requires screening of the small guide RNA as studies have shown that guides targeting a different region of RNA resulted in a different suppression efficiency, thus being more time-consuming.
Additionally, it was also shown that Cas13 could mediate RNA editing. It was shown that by fusing the dCas13 with an enzyme named adenosine deaminase acting on RNA (ADAR), the base substitution of adenosine to inosine (G) in RNA transcripts, could be performed [166]. This base substitution would reconstitute the mRNA sequence thus affecting translation.
Using these systems, miRNA could either be cleaved or edited. For example, in cases such as a point mutation occurring in a miRNA gene sequence, the mutation could be edited via dCas13. The repaired miRNA could resume its normal function as a single point change in the seed sequence would hamper miRNA function [81,168]. Additionally, Cas13 could also be used to down-regulate specific miRNA activity by designing the small guide sequence to be complementary towards the target miRNA. Currently, Cas13 has the potential to target miRNA but research still lacks in this field. This opens up a great opportunity for Cas13 research into miRNA targeting.

4.5. Utilization of CRISPR to Target miRNA: Practical Considerations

The previous section of this review has shown that arrays of CRISPR techniques are available for miRNA targeting. In view of this, some important considerations which could serve as a general guideline will be highlighted shortly.
Few strategies could be employed depending on desired outcomes. miRNA for example could be indirectly targeted by designing CRISPR against their activators, transcription factors responsible for driving their expression, as in the case of c-MYC [169]. This could repress the expression of targeted miRNA via transcriptional block. In contrast, suppressors could also be targeted in order to induce the expression of a targeted miRNA. However, targeting TFs would induce a global expression change as these TFs could also bind to other promoters containing its consensus binding sequence, thus it is not advisable to do this (Figure 6A).
Apparently, this could be mitigated by targeting a specific site by engineering sgRNA sequence directing dCas9 towards that particular site. This would not cause a change in global expression pattern as the dCas9 blocks TFs from actually binding onto their consensus sequence found only on the miRNA locus, thus achieving a specific miRNA knockdown. It should also be understood that these sequences are promoters/enhancers and that promoters could either be polycistronic or monocistronic [170,171]. If a particular miRNA locus is driven by only one promoter doing this is favorable, if it is polycistronic where a particular promoter can actually drive multiple miRNAs or even genes within close proximity to the miRNA, this technique becomes unfeasible.
In the case where miRNA overexpression is caused by the amplification of loci, the Cas9 nickase strategy could be employed. This could be achieved by designing two sgRNAs that complement flanking amplified sites, where it would result in “nicking” or removal of a fragment of targeted loci [172,173] (Figure 6B). Amplified fragments could be removed using this method. However, miRNA are short sequences of RNA and that only the seed sequences are responsible for complementing target mRNA therefore, removal of large loci is not advisable. Additionally, via homology-directed repair (HDR), a fragment with the “correct” sequence could be re-introduced to restore the “nicked” fragment of the genome. This is crucial as to not alter the sequence which may be crucial for other cellular gene activity.
Epigenetics plays an important role in the expression of miRNA, chromatin state brought about by histone deacetylase (HDAC) and acetyl transferase (HAT) plays a huge role in allowing TF accessibility onto a genomic region that drives the expression of that particular miRNA. Euchromatin state in heterochromatin can be induced by bringing the p300 close to the region of interest that could be in supercoiled state (heterochromatin) [174,175] (Figure 6C). p300 is an activator domain that acts as histone acetyl transferase, where it catalyzes the transfer of acetyl group onto histones. This process relaxes the DNA-Histone coil allowing the region targeted with p300 to be more accessible towards the TF/pol II complex which would result in an increase in transcriptional activity. This is especially helpful when applied to activate downregulated miRNA crucial for proper development. Additionally, if a particular miRNA is over-expressed, dCas9-KRAB is useful where it helps in forming heterochromatin on the highly accessible miRNA loci as explained earlier [150]. Therefore, it is crucial to identify the type of dysregulation occurring on the miRNA loci before proceeding with this method. One way of doing so is to identify enrichment of histone states via ChIP-qPCR (for a specific site) or ChIP-Seq (for global enrichment of histone states) [176].
An overabundance of CpG islands often led to gene silencing; using Tet-CD (Ten eleven translocationcatalytic domain) this could be alleviated via its demethylase activity. Tet-CD is a functional domain of a demethylase responsible for oxidizing the methyl on cytosine (Figure 6D). This helps in reducing DNA methylation which would promote gene activation. It was clearly shown that the abundance of CpG islands on miRNA loci would impair expression resulting in the downregulation of particular miRNAs crucial for proper development. Therefore, via CRISPR, Tet-CD fused to dCas9 could be used to specifically target Tet-CD onto the CpG rich sites allowing for demethylation of target sequence thus restoring miRNA expression adjacent to the targeted site [177,178]. CpG abundance across a particular miRNA genome can be identified by bisulfide sequencing which measures relative levels of methylation across loci [179].
A summary of the various CRISPR based techniques used to study miRNA dysregulation in cancer is depicted in Table 3. Table 3: Various CRISPR based techniques that are used to study miRNA function in cancer.

5. Conclusions: Challenges and Future Perspectives

It is clearly apparent that studying miRNA dysregulation in cancer is vital for us to understand cancer biology. A wide array of molecular biology techniques are readily accessible for the researcher to conduct the relevant experiments. Nevertheless, extreme care is needed to choose the correct molecular biology tool in order to answer the relevant research questions in regard to miRNA dysregulation in cancer. A general workflow is provided in Figure 7.
Most researchers will find it tempting to utilize CRISPR/Cas-based techniques for studying miRNA dysfunction due to their specificity and exquisiteness. However, there are certain issues and challenges that should be taken into account when employing these techniques in miRNA targeting. This is mainly due to the complexity of miRNA regulation. Points to consider include miRNA location, distribution and targets. This was exemplified by the miR-17-92 cluster which was implicated in many cancers [189]. However, the miR-17-92 is a cluster where different miRNAs reside proximally between one another; this will prove to be a challenge when using Cas9, as indels may induce mutation into the miRNA sequences resulting in either miRNA that is null or off-target miRNA. Additionally, clusters of miRNA are usually polycistronic, where a single promoter drives the expression of multiple miRNAs. Therefore, targeting a single miRNA in a cluster would be a challenge since this would affect adjacent miRNAs.
Micro RNA that resides within exonic sequence would also hamper targeting via CRISPR. Targeting these miRs may result in disruption of the splicing machinery leading to changes in the mRNA component of the host gene. This is a major hole in miR targeting using CRISPR. Even if there is an argument of using dCas9 for the transcriptional block instead of genome cutting considerations on how the mechanism induces silencing must be taken. Dead Cas9 blocks transcription elongation, targeting miRNA located in the host gene. Using this method is not advised as the experimenter must consider the simultaneous silencing of both target miR and host gene.
miRNA seed sequence targeting complexity of genes would also prove to be a challenge in direct targeting. In breast cancer (BC) it was found that miR-139-5p targets multiple genes involved in crucial BC oncogenesis [190]. Even with CRISPR/Cas-mediated suppression of miR, phenotypic changes would not be induced by a single gene expressional change. This requires further screening of the genes which would be time-consuming and costly. Furthermore, direct miRNA targeting is not as straightforward as targeting a different region of the miR transcript would result in a different suppression efficiency [191].
In conclusion, it can be said that CRISPR/Cas 9-based tools will gradually replace other conventional methods in studying miRNA dysregulation. Nevertheless, the relevance of more conventional techniques such as miRNA sponges, antagomirs and miRNA mimic cannot be negated altogether as these tools could prove to be extremely useful, especially when the specific miRNA is not pliable to CRISPR/Cas-based genetic manipulation.

Author Contributions

Conceptualization, E.J.M. and A.A.; Writing—Original Draft Preparation, E.J.M., M.Y.Y. and A.A. Writing—Review and Editing, Y.R., E.J.M., M.Y.Y., A.A., K.K.Z. and A.A.R.; Supervision, E.J.M., O.H. and N.M.Y. All authors have read and agreed to the published version of the manuscript.

Funding

Studies related to this topic were funded and supported by Fundamental Research Grant Scheme (FRGS): FRGS/1/2018/SKK08/USM/02/8 from Ministry of Higher Education (MOHE), Malaysia and Universiti Sains Malaysia Research Grant, USM–RUI: 1001/CIPPT/8012265.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lee, R.C.; Feinbaum, R.L.; Ambros, V. The C elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell 1993, 75, 843–854. [Google Scholar] [CrossRef]
  2. Lai, E.C. miRNAs: Whys and wherefores of miRNA-mediated regulation. Curr. Biol. 2005, 15, R458–R460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Lai, E.C. Two decades of miRNA biology: Lessons and challenges. RNA 2015, 21, 675–677. [Google Scholar] [CrossRef]
  4. Chipman, L.B.; Pasquinelli, A.E. miRNA targeting: Growing beyond the seed. Trends Genet. 2019, 35, 215–222. [Google Scholar] [CrossRef] [PubMed]
  5. Saliminejad, K.; Khorram Khorshid, H.R.; Soleymani Fard, S.; Ghaffari, S.H. An overview of microRNAs: Biology, functions, therapeutics, and analysis methods. J. Cell. Physiol. 2019, 234, 5451–5465. [Google Scholar] [CrossRef] [PubMed]
  6. Esteller, M. Non-coding RNAs in human disease. Nat. Rev. Genet. 2011, 12, 861–874. [Google Scholar] [CrossRef] [PubMed]
  7. Di Leva, G.; Croce, C.M. miRNA profiling of cancer. Curr. Opin. Genet. Dev. 2013, 23, 3–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Croce, C.M. Causes and consequences of microRNA dysregulation in cancer. Nat. Rev. Genet. 2009, 10, 704–714. [Google Scholar] [CrossRef] [PubMed]
  9. Zare, M.; Bastami, M.; Solali, S.; Alivand, M.R. Aberrant miRNA promoter methylation and EMT-involving miRNAs in breast cancer metastasis: Diagnosis and therapeutic implications. J. Cell Physiol. 2018, 233, 3729–3744. [Google Scholar] [CrossRef] [PubMed]
  10. Ryan, B.M. microRNAs in cancer susceptibility. Adv. Cancer Res. 2017, 135, 151–171. [Google Scholar]
  11. Markopoulos, G.S.; Roupakia, E.; Tokamani, M.; Chavdoula, E.; Hatziapostolou, M.; Polytarchou, C.; Marcu, K.B.; Papavassiliou, A.G.; Sandaltzopoulos, R.; Kolettas, E. A step-by-step microRNA guide to cancer development and metastasis. Cell. Oncol. 2017, 40, 303–339. [Google Scholar] [CrossRef]
  12. Lee, J.Y.; Kim, S.; Jeong, J.M.; Chung, J.-K.; Lee, M.C.; Lee, D.S. Development of a dual-luciferase reporter system for in vivo visualization of MicroRNA biogenesis and posttranscriptional regulation. J. Nucl. Med. 2008, 49, 285–294. [Google Scholar] [CrossRef] [Green Version]
  13. Fang, Y.; Zhou, Y.; Zhang, Y.; He, L.; Xue, C.; Cao, Y. Design of miRNA sponges for MDV-1 as a therapeutic strategy against lymphomas. Oncotarget 2018, 9, 3842–3852. [Google Scholar] [CrossRef] [Green Version]
  14. Jin, Y.; Chen, Z.; Liu, X.; Zhou, X. Evaluating the microRNA targeting sites by luciferase reporter gene assay. In MicroRNA Protocols; Springer: Berlin/Heidelberg, Germany, 2013; pp. 117–127. [Google Scholar]
  15. Chen, L.; Heikkinen, L.; Wang, C.; Yang, Y.; Sun, H.; Wong, G. Trends in the development of miRNA bioinformatics tools. Brief. Bioinform. 2019, 20, 1836–1852. [Google Scholar] [CrossRef] [Green Version]
  16. Ardekani, A.M.; Naeini, M.M. The Role of MicroRNAs in Human Diseases. Avicenna J. Med. Biotechnol. 2010, 2, 161–179. [Google Scholar] [PubMed]
  17. Rupaimoole, R.; Slack, F.J. MicroRNA therapeutics: Towards a new era for the management of cancer and other diseases. Nat. Rev. Drug Discov. 2017, 16, 203–222. [Google Scholar] [CrossRef] [PubMed]
  18. Agarwal, V.; Bell, G.W.; Nam, J.-W.; Bartel, D.P. Predicting effective microRNA target sites in mammalian mRNAs. eLife 2015, 4, e05005. [Google Scholar] [CrossRef]
  19. Wang, X. Improving microRNA target prediction by modeling with unambiguously identified microRNA-target pairs from CLIP-ligation studies. Bioinformatics 2016, 32, 1316–1322. [Google Scholar] [CrossRef] [PubMed]
  20. Chou, C.-H.; Shrestha, S.; Yang, C.-D.; Chang, N.-W.; Lin, Y.-L.; Liao, K.-W.; Huang, W.-C.; Sun, T.-H.; Tu, S.-J.; Lee, W.-H.; et al. miRTarBase update 2018: A resource for experimentally validated microRNA-target interactions. Nucleic Acids Res. 2018, 46, D296–D302. [Google Scholar] [CrossRef] [PubMed]
  21. Dweep, H.; Gretz, N. miRWalk2.0: A comprehensive atlas of microRNA-target interactions. Nat. Methods 2015, 12, 697. [Google Scholar] [CrossRef] [PubMed]
  22. Xie, C.; Chen, Y.-L.; Wang, D.-F.; Wang, Y.-L.; Zhang, T.-P.; Li, H.; Liang, F.; Zhao, Y.; Zhang, G.Y. SgRNA Expression of CRIPSR-Cas9 System Based on MiRNA Polycistrons as a Versatile Tool to Manipulate Multiple and Tissue-Specific Genome Editing. Sci. Rep. 2017, 19, 5795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Sun, P.; Liu, D.Z.; Jickling, G.C.; Sharp, F.R.; Yin, K.-J. MicroRNA-based therapeutics in central nervous system injuries. J. Cereb. Blood Flow Metab. Off. J. Int. Soc. Cereb. Blood Flow Metab. 2018, 38, 1125–1148. [Google Scholar] [CrossRef]
  24. Yi, B.; Larter, K.; Xi, Y. CRISPR/Cas9 System to Knockdown MicroRNA In Vitro and In Vivo. In Small Non-Coding RNAs: Methods and Protocols; Rederstorff, M., Ed.; Methods in Molecular Biology; Springer US: New York, NY, USA, 2021; pp. 133–139. ISBN 978-1-07-161386-3. [Google Scholar]
  25. Lataniotis, L.; Albrecht, A.; Kok, F.O.; Monfries, C.A.; Benedetti, L.; Lawson, N.D.; Hughes, S.M.; Steinhofel, K.; Mayr, M.; Zampetaki, A. CRISPR/Cas9 editing reveals novel mechanisms of clustered microRNA regulation and function. Sci. Rep. 2017, 7, 1–14. [Google Scholar]
  26. Zhou, J.; Deng, K.; Cheng, Y.; Zhong, Z.; Tian, L.; Tang, X.; Tang, A.; Zheng, X.; Zhang, T.; Qi, Y. CRISPR-Cas9 based genome editing reveals new insights into microRNA function and regulation in rice. Front. Plant Sci. 2017, 8, 1598. [Google Scholar] [CrossRef] [Green Version]
  27. Chaluvally-Raghavan, P.; Jeong, K.J.; Pradeep, S.; Silva, A.M.; Yu, S.; Liu, W.; Moss, T.; Rodriguez-Aguayo, C.; Zhang, D.; Ram, P.; et al. Direct upregulation of STAT3 by microRNA-551b-3p deregulates growth and metastasis of ovarian cancer. Cell Rep. 2016, 15, 1493–1504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Li, M.; Lee, K.F.; Lu, Y.; Clarke, I.; Shih, D.; Eberhart, C.; Collins, V.P.; Van Meter, T.; Picard, D.; Zhou, L.; et al. Frequent amplification of a chr19q13.41 microRNA polycistron (C19MC) in aggressive primitive neuro-ectodermal brain tumors. Cancer Cell 2009, 16, 533–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Spence, T.; Sin-Chan, P.; Picard, D.; Barszczyk, M.; Hoss, K.; Lu, M.; Kim, S.-K.; Ra, Y.-S.; Nakamura, H.; Fangusaro, J.; et al. CNS-PNETs with C19MC amplification and/or LIN28 expression comprise a distinct histogenetic diagnostic and therapeutic entity. Acta Neuropathol. 2014, 128, 291–303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Hayashita, Y.; Osada, H.; Tatematsu, Y.; Yamada, H.; Yanagisawa, K.; Tomida, S.; Yatabe, Y.; Kawahara, K.; Sekido, Y.; Takahashi, T. A Polycistronic MicroRNA Cluster, miR-17-92, Is Overexpressed in Human Lung Cancers and Enhances Cell Proliferation. Cancer Res. 2005, 65, 9628–9632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Mi, S.; Li, Z.; Chen, P.; He, C.; Cao, D.; Elkahloun, A.; Lu, J.; Pelloso, L.A.; Wunderlich, M.; Huang, H.; et al. Aberrant overexpression and function of the miR-17-92 cluster in MLL-rearranged acute leukemia. Proc. Natl. Acad. Sci. USA 2010, 107, 3710–3715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. An, J.; Pan, Y.; Yan, Z.; Li, W.; Cui, J.; Yuan, J.; Tian, L.; Xing, R.; Lu, Y. MiR-23a in amplified 19p13.13 loci targets metallothionein 2A and promotes growth in gastric cancer cells. J. Cell. Biochem. 2013, 114, 2160–2169. [Google Scholar] [CrossRef]
  33. Zhang, X.; Wu, M.; Chong, Q.-Y.; Zhang, W.; Qian, P.; Yan, H.; Qian, W.; Zhang, M.; Lobie, P.E.; Zhu, T. Amplification of hsa-miR-191/425 locus promotes breast cancer proliferation and metastasis by targeting DICER1. Carcinogenesis 2018, 39, 1506–1516. [Google Scholar] [CrossRef]
  34. O’Donoghue, P.; Prat, L.; Kucklick, M.; Schäfer, J.G.; Riedel, K.; Rinehart, J.; Söll, D.; Heinemann, I.U. Reducing the genetic code induces massive rearrangement of the proteome. Proc. Natl. Acad. Sci. USA 2014, 111, 17206–17211. [Google Scholar] [CrossRef] [Green Version]
  35. Drummond, D.A.; Wilke, C.O. The evolutionary consequences of erroneous protein synthesis. Nat. Rev. Genet. 2009, 10, 715–724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Zhang, Z.; Wong, G.K.-S.; Yu, J. Protein Coding. In Protein Coding; John Wiley & Sons, Ltd: Chichester, UK, 2013; ISBN 978-0-470-01617-6. [Google Scholar]
  37. Elli, F.M.; Ghirardello, S.; Giavoli, C.; Gangi, S.; Dioni, L.; Crippa, M.; Finelli, P.; Bergamaschi, S.; Mosca, F.; Spada, A. A new structural rearrangement associated to Wolfram syndrome in a child with a partial phenotype. Gene 2012, 509, 168–172. [Google Scholar] [CrossRef] [PubMed]
  38. Lu, X.; Cen, Z.; Xie, F.; Ouyang, Z.; Zhang, B.; Zhao, G.; Luo, W. Genetic analysis of SPG4 and SPG3A genes in a cohort of Chinese patients with hereditary spastic paraplegia. J. Neurol. Sci. 2014, 347, 368–371. [Google Scholar] [CrossRef] [PubMed]
  39. Ou, S.-H.I.; Azada, M.; Hsiang, D.J.; Herman, J.M.; Kain, T.S.; Siwak-Tapp, C.; Casey, C.; He, J.; Ali, S.M.; Klempner, S.J. Next-generation sequencing reveals a Novel NSCLC ALK F1174V mutation and confirms ALK G1202R mutation confers high-level resistance to alectinib (CH5424802/RO5424802) in ALK-rearranged NSCLC patients who progressed on crizotinib. J. Thorac. Oncol. 2014, 9, 549–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Abe, A.; Yamamoto, Y.; Katsumi, A.; Okamoto, A.; Tokuda, M.; Inaguma, Y.; Yamamoto, K.; Yanada, M.; Kanie, T.; Tomita, A. Rearrangement of VPS13B, a causative gene of Cohen syndrome, in a case of RUNX1–RUNX1T1 leukemia with t (8; 12; 21). Int. J. Hematol. 2018, 108, 208–212. [Google Scholar] [CrossRef] [PubMed]
  41. Porkka, K.P.; Ogg, E.-L.; Saramäki, O.R.; Vessella, R.L.; Pukkila, H.; Lähdesmäki, H.; Weerden, W.M.; van Wolf, M.; Kallioniemi, O.P.; Jenster, G.; et al. The miR-15a-miR-16-1 locus is homozygously deleted in a subset of prostate cancers. Genes Chromosomes Cancer 2011, 50, 499–509. [Google Scholar] [CrossRef]
  42. Trissal, M.C.; Wong, T.N.; Yao, J.-C.; Ramaswamy, R.; Kuo, I.; Baty, J.; Sun, Y.; Jih, G.; Parikh, N.; Berrien-Elliott, M.M.; et al. MIR142 Loss-of-Function Mutations Derepress ASH1L to Increase HOXA Gene Expression and Promote Leukemogenesis. Cancer Res. 2018, 78, 3510–3521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Urbanek-Trzeciak, M.O.; Galka-Marciniak, P.; Nawrocka, P.M.; Kowal, E.; Szwec, S.; Giefing, M.; Kozlowski, P. Pan-cancer analysis of somatic mutations in miRNA genes. EBioMedicine 2020, 61, 103051. [Google Scholar] [CrossRef] [PubMed]
  44. Kwanhian, W.; Lenze, D.; Alles, J.; Motsch, N.; Barth, S.; Döll, C.; Imig, J.; Hummel, M.; Tinguely, M.; Trivedi, P.; et al. MicroRNA-142 is mutated in about 20% of diffuse large B-cell lymphoma. Cancer Med. 2012, 1, 141–155. [Google Scholar] [CrossRef] [PubMed]
  45. Meng, F.; Wehbe-Janek, H.; Henson, R.; Smith, H.; Patel, T. Epigenetic regulation of microRNA-370 by interleukin-6 in malignant human cholangiocytes. Oncogene 2008, 27, 378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Iorio, M.V.; Piovan, C.; Croce, C.M. Interplay between microRNAs and the epigenetic machinery: An intricate network. Biochim. Biophys. Acta BBA-Gene Regul. Mech. 2010, 1799, 694–701. [Google Scholar] [CrossRef] [PubMed]
  47. Song, K.; Han, C.; Zhang, J.; Lu, D.; Dash, S.; Feitelson, M.; Lim, K.; Wu, T. Epigenetic regulation of miR-122 by PPARγ and hepatitis B virus X protein in hepatocellular carcinoma cells. Hepatology 2013, 58, 1681–1692. [Google Scholar] [CrossRef] [Green Version]
  48. Bourassa, M.W.; Ratan, R.R. The interplay between microRNAs and histone deacetylases in neurological diseases. Neurochem. Int. 2014, 77, 33–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Tomasetti, M.; Gaetani, S.; Monaco, F.; Neuzil, J.; Santarelli, L. Epigenetic Regulation of miRNA Expression in Malignant Mesothelioma: miRNAs as Biomarkers of Early Diagnosis and Therapy. Front. Oncol. 2019, 9, 1293. [Google Scholar] [CrossRef] [PubMed]
  50. Li, G.; Tian, Y.; Zhu, W.-G. The Roles of Histone Deacetylases and Their Inhibitors in Cancer Therapy. Front. Cell Dev. Biol. 2020, 8, 1004. [Google Scholar] [CrossRef] [PubMed]
  51. Hsieh, T.-H.; Hsu, C.-Y.; Tsai, C.-F.; Long, C.-Y.; Wu, C.-H.; Wu, D.-C.; Lee, J.-N.; Chang, W.-C.; Tsai, E.-M. HDAC Inhibitors Target HDAC5, Upregulate MicroRNA-125a-5p, and Induce Apoptosis in Breast Cancer Cells. Mol. Ther. 2015, 23, 656–666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Li, C.H.; To, K.; Tong, J.H.; Xiao, Z.; Xia, T.; Lai, P.B.S.; Chow, S.C.; Zhu, Y.; Chan, S.L.; Marquez, V.E.; et al. Enhancer of Zeste Homolog 2 Silences MicroRNA-218 in Human Pancreatic Ductal Adenocarcinoma Cells by Inducing Formation of Heterochromatin. Gastroenterology 2013, 144, 1086–1097. [Google Scholar] [CrossRef] [PubMed]
  53. Dohi, O.; Yasui, K.; Gen, Y.; Takada, H.; Endo, M.; Tsuji, K.; Konishi, C.; Yamada, N.; Mitsuyoshi, H.; Yagi, N. Epigenetic silencing of miR-335 and its host gene MEST in hepatocellular carcinoma. Int. J. Oncol. 2013, 42, 411–418. [Google Scholar] [CrossRef] [Green Version]
  54. Srivastava, A.; Goldberger, H.; Dimtchev, A.; Ramalinga, M.; Chijioke, J.; Marian, C.; Oermann, E.K.; Uhm, S.; Kim, J.S.; Chen, L.N. MicroRNA profiling in prostate cancer-the diagnostic potential of urinary miR-205 and miR-214. PLoS ONE 2013, 8, e76994. [Google Scholar] [CrossRef] [Green Version]
  55. Gandellini, P.; Giannoni, E.; Casamichele, A.; Taddei, M.L.; Callari, M.; Piovan, C.; Valdagni, R.; Pierotti, M.A.; Zaffaroni, N.; Chiarugi, P. miR-205 hinders the malignant interplay between prostate cancer cells and associated fibroblasts. Antioxid. Redox Signal 2014, 20, 1045–1059. [Google Scholar] [CrossRef] [Green Version]
  56. Hulf, T.; Sibbritt, T.; Wiklund, E.D.; Patterson, K.; Song, J.Z.; Stirzaker, C.; Qu, W.; Nair, S.; Horvath, L.G.; Armstrong, N.J.; et al. Epigenetic-induced repression of microRNA-205 is associated with MED1 activation and a poorer prognosis in localized prostate cancer. Oncogene 2013, 32, 2891–2899. [Google Scholar] [CrossRef] [Green Version]
  57. Babu, A.; Verma, R.S. Chromosome structure: Euchromatin and heterochromatin. In International Review of Cytology; Elsevier: Amsterdam, The Netherlands, 1987; Volume 108, pp. 1–60. [Google Scholar]
  58. Wang, X.; Wang, J.; Ma, H.; Zhang, J.; Zhou, X. Downregulation of miR-195 correlates with lymph node metastasis and poor prognosis in colorectal cancer. Med. Oncol. 2012, 29, 919–927. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, L.; Chen, L.; Xu, Y.; Li, R.; Du, X. microRNA-195 promotes apoptosis and suppresses tumorigenicity of human colorectal cancer cells. Biochem. Biophys. Res. Commun. 2010, 400, 236–240. [Google Scholar] [CrossRef]
  60. Menigatti, M.; Staiano, T.; Manser, C.N.; Bauerfeind, P.; Komljenovic, A.; Robinson, M.; Jiricny, J.; Buffoli, F.; Marra, G. Epigenetic silencing of monoallelically methylated miRNA loci in precancerous colorectal lesions. Oncogenesis 2013, 2, e56. [Google Scholar] [CrossRef] [Green Version]
  61. Domingo-Fernandez, R.; Watters, K.; Piskareva, O.; Stallings, R.L.; Bray, I. The role of genetic and epigenetic alterations in neuroblastoma disease pathogenesis. Pediatr. Surg. Int. 2013, 29, 101–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Ram Kumar, R.M.; Schor, N.F. Methylation of DNA and chromatin as a mechanism of oncogenesis and therapeutic target in neuroblastoma. Oncotarget 2018, 9, 22184–22193. [Google Scholar] [CrossRef] [Green Version]
  63. Das, S.; Bryan, K.; Buckley, P.G.; Piskareva, O.; Bray, I.M.; Foley, N.; Ryan, J.; Lynch, J.; Creevey, L.; Fay, J.; et al. Modulation of neuroblastoma disease pathogenesis by an extensive network of epigenetically regulated microRNAs. Oncogene 2013, 32, 2927–2936. [Google Scholar] [CrossRef] [Green Version]
  64. Harbison, C.T.; Gordon, D.B.; Lee, T.I.; Rinaldi, N.J.; Macisaac, K.D.; Danford, T.W.; Hannett, N.M.; Tagne, J.-B.; Reynolds, D.B.; Yoo, J. Transcriptional regulatory code of a eukaryotic genome. Nature 2004, 431, 99. [Google Scholar] [CrossRef] [PubMed]
  65. Wingender, E.; Dietze, P.; Karas, H.; Knüppel, R. TRANSFAC: A database on transcription factors and their DNA binding sites. Nucleic Acids Res. 1996, 24, 238–241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Cawley, S.; Bekiranov, S.; Ng, H.H.; Kapranov, P.; Sekinger, E.A.; Kampa, D.; Piccolboni, A.; Sementchenko, V.; Cheng, J.; Williams, A.J. Unbiased mapping of transcription factor binding sites along human chromosomes 21 and 22 points to widespread regulation of noncoding RNAs. Cell 2004, 116, 499–509. [Google Scholar] [CrossRef] [Green Version]
  67. Stadler, M.B.; Murr, R.; Burger, L.; Ivanek, R.; Lienert, F.; Schöler, A.; van Nimwegen, E.; Wirbelauer, C.; Oakeley, E.J.; Gaidatzis, D. DNA-binding factors shape the mouse methylome at distal regulatory regions. Nature 2011, 480, 490–495. [Google Scholar] [CrossRef]
  68. Spitz, F.; Furlong, E.E. Transcription factors: From enhancer binding to developmental control. Nat. Rev. Genet. 2012, 13, 613. [Google Scholar] [CrossRef] [PubMed]
  69. Pei, L.; He, X.; Li, S.; Sun, R.; Xiang, Q.; Ren, G.; Xiang, T. KRAB zinc-finger protein 382 regulates epithelial-mesenchymal transition and functions as a tumor suppressor, but is silenced by CpG methylation in gastric cancer. Int. J. Oncol. 2018, 53, 961–972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Puisieux, A.; Brabletz, T.; Caramel, J. Oncogenic roles of EMT-inducing transcription factors. Nat. Cell Biol. 2014, 16, 488. [Google Scholar] [CrossRef]
  71. Shu, X.-Z.; Zhang, L.-N.; Zhang, R.; Zhang, C.-J.; He, H.-P.; Zhou, H.; Wang, N.; Zhang, T.-C. Histone acetyltransferase p300 promotes MRTF-A-mediates transactivation of VE-cadherin gene in human umbilical vein endothelial cells. Gene 2015, 563, 17–23. [Google Scholar] [CrossRef] [PubMed]
  72. Quinn, S.R.; Mangan, N.E.; Caffrey, B.E.; Gantier, M.P.; Williams, B.R.; Hertzog, P.J.; McCoy, C.E.; O’Neill, L.A. The role of Ets2 transcription factor in the induction of microRNA-155 (miR-155) by lipopolysaccharide and its targeting by interleukin-10. J. Biol. Chem. 2014, 289, 4316–4325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Liu, H.; Wang, B.; Lin, J.; Zhao, L. microRNA-29b: An emerging player in human cancer. Asian Pac. J. Cancer Prev. 2014, 15, 9059–9064. [Google Scholar] [CrossRef] [Green Version]
  74. Zhang, K.; Cai, H.-X.; Gao, S.; Yang, G.-L.; Deng, H.-T.; Xu, G.-C.; Han, J.; Zhang, Q.-Z.; Li, L.-Y. TNFSF15 suppresses VEGF production in endothelial cells by stimulating miR-29b expression via activation of JNK-GATA3 signals. Oncotarget 2016, 7, 69436–69449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Wang, S.-H.; Zhang, W.-J.; Wu, X.-C.; Zhang, M.-D.; Weng, M.-Z.; Zhou, D.; Wang, J.-D.; Quan, Z.-W. Long non-coding RNA Malat1 promotes gallbladder cancer development by acting as a molecular sponge to regulate miR-206. Oncotarget 2016, 7, 37857. [Google Scholar] [CrossRef] [Green Version]
  76. Han, C.; Shen, J.K.; Hornicek, F.J.; Kan, Q.; Duan, Z. Regulation of microRNA-1 (miR-1) expression in human cancer. Biochim. Biophys. Acta BBA-Gene Regul. Mech. 2017, 1860, 227–232. [Google Scholar] [CrossRef] [PubMed]
  77. Lango-Chavarría, M.; Chimal-Ramírez, G.K.; Ruiz-Tachiquín, M.E.; Espinoza-Sánchez, N.A.; Suárez-Arriaga, M.C.; Fuentes-Pananá, E.M. A 22q11.2 amplification in the region encoding microRNA-650 correlates with the epithelial to mesenchymal transition in breast cancer primary cultures of Mexican patients. Int. J. Oncol. 2017, 50, 432–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Hirata, Y.; Murai, N.; Yanaihara, N.; Saito, M.; Saito, M.; Urashima, M.; Murakami, Y.; Matsufuji, S.; Okamoto, A. MicroRNA-21 is a candidate driver gene for 17q23-25 amplification in ovarian clear cell carcinoma. BMC Cancer 2014, 14, 799. [Google Scholar] [CrossRef] [Green Version]
  79. Bhagirath, D.; Yang, T.L.; Tabatabai, Z.L.; Shahryari, V.; Majid, S.; Dahiya, R.; Tanaka, Y.; Saini, S. Role of a novel race-related tumor suppressor microRNA located in frequently deleted chromosomal locus 8p21 in prostate cancer progression. Carcinogenesis 2019, 40, 633–642. [Google Scholar] [CrossRef] [PubMed]
  80. Chen, C.; Pan, Y.; Bai, L.; Chen, H.; Duan, Z.; Si, Q.; Zhu, R.; Chuang, T.-H.; Luo, Y. MicroRNA-3613-3p functions as a tumor suppressor and represents a novel therapeutic target in breast cancer. Breast Cancer Res. 2021, 23, 12. [Google Scholar] [CrossRef]
  81. Galka-Marciniak, P.; Urbanek-Trzeciak, M.O.; Nawrocka, P.M.; Dutkiewicz, A.; Giefing, M.; Lewandowska, M.A.; Kozlowski, P. Somatic Mutations in miRNA Genes in Lung Cancer—Potential Functional Consequences of Non-Coding Sequence Variants. Cancers 2019, 11, 793. [Google Scholar] [CrossRef] [Green Version]
  82. Li, Y.; He, Q.; Wen, X.; Hong, X.; Yang, X.; Tang, X.; Zhang, P.; Lei, Y.; Sun, Y.; Zhang, J.; et al. EZH2-DNMT1-mediated epigenetic silencing of miR-142-3p promotes metastasis through targeting ZEB2 in nasopharyngeal carcinoma. Cell Death Differ. 2019, 26, 1089–1106. [Google Scholar] [CrossRef]
  83. Bhatia, V.; Yadav, A.; Tiwari, R.; Nigam, S.; Goel, S.; Carskadon, S.; Gupta, N.; Goel, A.; Palanisamy, N.; Ateeq, B. Epigenetic Silencing of miRNA-338-5p and miRNA-421 Drives SPINK1-Positive Prostate Cancer. Clin. Cancer Res. 2019, 25, 2755–2768. [Google Scholar] [CrossRef] [Green Version]
  84. Li, S.-G.; Shi, Q.-W.; Yuan, L.; Qin, L.; Wang, Y.; Miao, Y.-Q.; Chen, Z.; Ling, C.-Q.; Qin, W. C-Myc-dependent repression of two oncogenic miRNA clusters contributes to triptolide-induced cell death in hepatocellular carcinoma cells. J. Exp. Clin. Cancer Res. 2018, 37, 51. [Google Scholar] [CrossRef]
  85. Wang, B.; Hsu, S.; Wang, X.; Kutay, H.; Bid, H.K.; Yu, J.; Ganju, R.K.; Jacob, S.T.; Yuneva, M.; Ghoshal, K. Reciprocal regulation of microRNA-122 and c-Myc in hepatocellular cancer: Role of E2F1 and transcription factor dimerization partner 2. Hepatology 2014, 59, 555–566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. La, T.; Liu, G.Z.; Farrelly, M.; Cole, N.; Feng, Y.C.; Zhang, Y.Y.; Sherwin, S.K.; Yari, H.; Tabatabaee, H.; Yan, X.G.; et al. A p53-Responsive miRNA Network Promotes Cancer Cell Quiescence. Cancer Res. 2018, 78, 6666–6679. [Google Scholar] [CrossRef] [Green Version]
  87. Bak, R.O.; Mikkelsen, J.G. miRNA sponges: Soaking up miRNAs for regulation of gene expression. Wiley Interdiscip. Rev. RNA 2014, 5, 317–333. [Google Scholar] [CrossRef]
  88. Wang, P.; Zhi, H.; Zhang, Y.; Liu, Y.; Zhang, J.; Gao, Y.; Guo, M.; Ning, S.; Li, X. miRSponge: A manually curated database for experimentally supported miRNA sponges and ceRNAs. Database 2015, 2015, bav098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Chen, L.; Zhang, S.; Wu, J.; Cui, J.; Zhong, L.; Zeng, L.; Ge, S. circRNA_100290 plays a role in oral cancer by functioning as a sponge of the miR-29 family. Oncogene 2017, 36, 4551. [Google Scholar] [CrossRef]
  90. Van Peer, G.; Mets, E.; Claeys, S.; De Punt, I.; Lefever, S.; Ongenaert, M.; Rondou, P.; Speleman, F.; Mestdagh, P.; Vandesompele, J. A high-throughput 3’UTR reporter screening identifies microRNA interactomes of cancer genes. PLoS ONE 2018, 13, e0194017. [Google Scholar] [CrossRef] [PubMed]
  91. Pan, Y.; Li, C.; Chen, J.; Zhang, K.; Chu, X.; Wang, R.; Chen, L. The emerging roles of long noncoding RNA ROR (lincRNA-ROR) and its possible mechanisms in human cancers. Cell. Physiol. Biochem. 2016, 40, 219–229. [Google Scholar] [CrossRef] [PubMed]
  92. Zhou, X.; Gao, Q.; Wang, J.; Zhang, X.; Liu, K.; Duan, Z. Linc-RNA-RoR acts as a “sponge” against mediation of the differentiation of endometrial cancer stem cells by microRNA-145. Gynecol. Oncol. 2014, 133, 333–339. [Google Scholar] [CrossRef]
  93. Bofill-De Ros, X.; Santos, M.; Vila-Casadesús, M.; Villanueva, E.; Andreu, N.; Dierssen, M.; Fillat, C. Genome-wide miR-155 and miR-802 target gene identification in the hippocampus of Ts65Dn Down syndrome mouse model by miRNA sponges. BMC Genom. 2015, 16, 907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Yang, X.; Cheng, Y.; Li, P.; Tao, J.; Deng, X.; Zhang, X.; Gu, M.; Lu, Q.; Yin, C. A lentiviral sponge for miRNA-21 diminishes aerobic glycolysis in bladder cancer T24 cells via the PTEN/PI3K/AKT/mTOR axis. Tumor Biol. 2015, 36, 383–391. [Google Scholar] [CrossRef]
  95. Wang, X. Composition of seed sequence is a major determinant of microRNA targeting patterns. Bioinformatics 2014, 30, 1377–1383. [Google Scholar] [CrossRef] [Green Version]
  96. Ebert, M.S.; Sharp, P.A. MicroRNA sponges: Progress and possibilities. RNA 2010, 16, 2043–2050. [Google Scholar] [CrossRef] [Green Version]
  97. Ebert, M.S.; Sharp, P.A. Emerging Roles for Natural MicroRNA Sponges. Curr. Biol. 2010, 20, R858–R861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Dias, N.; Stein, C.A. Antisense Oligonucleotides: Basic Concepts and Mechanisms. Mol. Cancer Ther. 2002, 1, 347–355. [Google Scholar] [PubMed]
  99. Zhou, S.; Wang, Y.; Meng, Y.; Xiao, C.; Liu, Z.; Brohawn, P.; Higgs, B.W.; Jallal, B.; Jia, Q.; Qu, B. In Vivo Therapeutic Success of MicroRNA-155 Antagomir in a Mouse Model of Lupus Alveolar Hemorrhage. Arthritis Rheumatol. 2016, 68, 953–964. [Google Scholar] [CrossRef]
  100. Dorrance, A.M.; Neviani, P.; Ferenchak, G.J.; Huang, X.; Nicolet, D.; Maharry, K.S.; Ozer, H.G.; Hoellarbauer, P.; Khalife, J.; Hill, E.B. Targeting leukemia stem cells in vivo with antagomiR-126 nanoparticles in acute myeloid leukemia. Leukemia 2015, 29, 2143. [Google Scholar] [CrossRef] [Green Version]
  101. Baker, J.; Vuppusetty, C.; Ito, K.; Barnes, P.J. Antagomir Of Microrna-34a Rescues Cellular Senescence In Bronchial Epithelial Cells Of COPD Patients. In D108 Cell Fate Decisions: Senescence, Repair, and Regeneration; American Thoracic Society: Oxford, UK, 2017; p. A7490. [Google Scholar]
  102. Yu, C.; Qian, L.; Uttamchandani, M.; Li, L.; Yao, S.Q. Single-vehicular delivery of antagomir and small molecules to inhibit miR-122 function in hepatocellular carcinoma cells by using “smart” mesoporous silica nanoparticles. Angew. Chem. Int. Ed. 2015, 54, 10574–10578. [Google Scholar] [CrossRef] [PubMed]
  103. Xu, L.-J.; Ouyang, Y.-B.; Xiong, X.; Stary, C.M.; Giffard, R.G. Post-stroke treatment with miR-181 antagomir reduces injury and improves long-term behavioral recovery in mice after focal cerebral ischemia. Exp. Neurol. 2015, 264, 1–7. [Google Scholar] [CrossRef] [Green Version]
  104. Brunschweiger, A.; Gebert, L.F.; Lucic, M.; Pradère, U.; Jahns, H.; Berk, C.; Hunziker, J.; Hall, J. Site-specific conjugation of drug-like fragments to an antimiR scaffold as a strategy to target miRNAs inside RISC. Chem. Commun. 2016, 52, 156–159. [Google Scholar] [CrossRef] [Green Version]
  105. Adams, B.D.; Parsons, C.; Walker, L.; Zhang, W.C.; Slack, F.J. Targeting noncoding RNAs in disease. J. Clin. Investig. 2017, 127, 761–771. [Google Scholar] [CrossRef] [PubMed]
  106. Cheng, C.J.; Bahal, R.; Babar, I.A.; Pincus, Z.; Barrera, F.; Liu, C.; Svoronos, A.; Braddock, D.T.; Glazer, P.M.; Engelman, D.M. MicroRNA silencing for cancer therapy targeted to the tumour microenvironment. Nature 2015, 518, 107. [Google Scholar] [CrossRef] [Green Version]
  107. Sun, J.-Y.; Xiao, W.-Z.; Wang, F.; Wang, Y.-Q.; Zhu, Y.-H.; Wu, Y.-F.; Miao, Z.-L.; Lin, Y.-C. MicroRNA-320 inhibits cell proliferation in glioma by targeting E2F1. Mol. Med. Rep. 2015, 12, 2355–2359. [Google Scholar] [CrossRef] [Green Version]
  108. Jia, Z.; Wang, K.; Wang, G.; Zhang, A.; Pu, P. MiR-30a-5p Antisense Oligonucleotide Suppresses Glioma Cell Growth by Targeting SEPT7. PLoS ONE 2013, 8, e55008. [Google Scholar] [CrossRef]
  109. Li, Z.; Chen, P.; Su, R.; Li, Y.; Hu, C.; Wang, Y.; Arnovitz, S.; He, M.; Gurbuxani, S.; Zuo, Z. Overexpression and knockout of miR-126 both promote leukemogenesis through targeting distinct gene signaling. Blood 2015, 126, 3667. [Google Scholar]
  110. Brewer, C.J.; Cook, G.J.; Zhang, B.; Qi, J.; Hua, W.-K.; Cai, Q.; Carnahan, E.; Marom, A.; Li, L.; Kortylewski, M. MiR-126 Promotes Leukemogenesis in Inv (16) Acute Myeloid Leukemia. Am. Soc. Hematol. 2017, 130, 301. [Google Scholar]
  111. Sharma, V.K.; Sharma, R.K.; Singh, S.K. Antisense oligonucleotides: Modifications and clinical trials. MedChemComm 2014, 5, 1454–1471. [Google Scholar] [CrossRef]
  112. Dowdy, S.F. Overcoming cellular barriers for RNA therapeutics. Nat. Biotechnol. 2017, 35, 222. [Google Scholar] [CrossRef] [PubMed]
  113. Braasch, D.A.; Corey, D.R. Locked nucleic acid (LNA): Fine-tuning the recognition of DNA and RNA. Chem. Biol. 2001, 8, 1–7. [Google Scholar] [CrossRef] [Green Version]
  114. Kurreck, J. Antisense technologies: Improvement through novel chemical modifications. Eur. J. Biochem. 2003, 270, 1628–1644. [Google Scholar] [CrossRef]
  115. Deleavey, G.F.; Damha, M.J. Designing Chemically Modified Oligonucleotides for Targeted Gene Silencing. Chem. Biol. 2012, 19, 937–954. [Google Scholar] [CrossRef] [Green Version]
  116. Stenvang, J.; Petri, A.; Lindow, M.; Obad, S.; Kauppinen, S. Inhibition of microRNA function by antimiR oligonucleotides. Silence 2012, 3, 1. [Google Scholar] [CrossRef] [Green Version]
  117. Van Rooij, E.; Kauppinen, S. Development of microRNA therapeutics is coming of age. EMBO Mol. Med. 2014, 6, 851–864. [Google Scholar] [CrossRef]
  118. Reid, G.; Kao, S.C.; Pavlakis, N.; Brahmbhatt, H.; MacDiarmid, J.; Clarke, S.; Boyer, M.; van Zandwijk, N. Clinical development of TargomiRs, a miRNA mimic-based treatment for patients with recurrent thoracic cancer. Epigenomics 2016, 8, 1079–1085. [Google Scholar] [CrossRef] [Green Version]
  119. Cutrona, G.; Matis, S.; Colombo, M.; Massucco, C.; Baio, G.; Valdora, F.; Emionite, L.; Fabris, S.; Recchia, A.G.; Gentile, M.; et al. Effects of miRNA-15 and miRNA-16 expression replacement in chronic lymphocytic leukemia: Implication for therapy. Leukemia 2017, 31, 1894–1904. [Google Scholar] [CrossRef] [Green Version]
  120. Yun, S.; Kim, W.K.; Kwon, Y.; Jang, M.; Bauer, S.; Kim, H. Survivin is a novel transcription regulator of KIT and is downregulated by miRNA-494 in gastrointestinal stromal tumors. Int. J. Cancer 2018, 142, 2080–2093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Shibayama, Y.; Kondo, T.; Ohya, H.; Fujisawa, S.-I.; Teshima, T.; Iseki, K. Upregulation of microRNA-126-5p is associated with drug resistance to cytarabine and poor prognosis in AML patients. Oncol. Rep. 2015, 33, 2176–2182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Søkilde, R.; Newie, I.; Persson, H.; Borg, A.; Rovira, C. Passenger strand loading in overexpression experiments using microRNA mimics. RNA Biol. 2015, 12, 787–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Jin, H.Y.; Gonzalez-Martin, A.; Miletic, A.V.; Lai, M.; Knight, S.; Sabouri-Ghomi, M.; Head, S.R.; Macauley, M.S.; Rickert, R.C.; Xiao, C. Transfection of microRNA Mimics Should Be Used with Caution. Front. Genet. 2015, 6. [Google Scholar] [CrossRef] [Green Version]
  124. Liu, X.; Abraham, J.M.; Cheng, Y.; Wang, Z.; Wang, Z.; Zhang, G.; Ashktorab, H.; Smoot, D.T.; Cole, R.N.; Boronina, T.N.; et al. Synthetic Circular RNA Functions as a miR-21 Sponge to Suppress Gastric Carcinoma Cell Proliferation. Mol. Ther. Nucleic Acids 2018, 13, 312–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Wang, Z.; Ma, K.; Cheng, Y.; Abraham, J.M.; Liu, X.; Ke, X.; Wang, Z.; Meltzer, S.J. Synthetic circular multi-miR sponge simultaneously inhibits miR-21 and miR-93 in esophageal carcinoma. Lab. Investig. 2019, 99, 1442–1453. [Google Scholar] [CrossRef] [PubMed]
  126. Shu, Y.; Wu, K.; Zeng, Z.; Huang, S.; Ji, X.; Yuan, C.; Zhang, L.; Liu, W.; Huang, B.; Feng, Y.; et al. A Simplified System to Express Circularized Inhibitors of miRNA for Stable and Potent Suppression of miRNA Functions. Mol. Ther. Nucleic Acids 2018, 13, 556–567. [Google Scholar] [CrossRef] [Green Version]
  127. Xu, W.; Song, C.; Wang, X.; Li, Y.; Bai, X.; Liang, X.; Wu, J.; Liu, J. Downregulation of miR-155-5p enhances the anti-tumor effect of cetuximab on triple-negative breast cancer cells via inducing cell apoptosis and pyroptosis. Aging 2021, 13, 228–240. [Google Scholar] [CrossRef] [PubMed]
  128. Li, R.; Hu, Z.; Wang, Z.; Zhu, T.; Wang, G.; Gao, B.; Wang, J.; Deng, X. miR-125a-5p promotes gastric cancer growth and invasion by regulating the Hippo pathway. J. Clin. Lab. Anal. 2021, 28, e24078. [Google Scholar] [CrossRef]
  129. Kardani, A.; Yaghoobi, H.; Alibakhshi, A.; Khatami, M. Inhibition of miR-155 in MCF-7 breast cancer cell line by gold nanoparticles functionalized with antagomir and AS1411 aptamer. J. Cell. Physiol. 2020, 235, 6887–6895. [Google Scholar] [CrossRef]
  130. Ueda, S.; Takanashi, M.; Sudo, K.; Kanekura, K.; Kuroda, M. miR-27a ameliorates chemoresistance of breast cancer cells by disruption of reactive oxygen species homeostasis and impairment of autophagy. Lab. Investig. 2020, 100, 863–873. [Google Scholar] [CrossRef]
  131. Fang, Z.H.; Wang, S.L.; Zhao, J.T.; Lin, Z.J.; Chen, L.Y.; Su, R.; Xie, S.T.; Carter, B.Z.; Xu, B. miR-150 exerts antileukemia activity in vitro and in vivo through regulating genes in multiple pathways. Cell Death Dis. 2016, 7, e2371. [Google Scholar] [CrossRef]
  132. Nuñez, J.K.; Lee, A.S.Y.; Engelman, A.; Doudna, J.A. Integrase-mediated spacer acquisition during CRISPR–Cas adaptive immunity. Nature 2015, 519, 193–198. [Google Scholar] [CrossRef] [Green Version]
  133. Bhaya, D.; Davison, M.; Barrangou, R. CRISPR-Cas systems in bacteria and archaea: Versatile small RNAs for adaptive defense and regulation. Annu. Rev. Genet. 2011, 45, 273–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Marraffini, L.A. CRISPR-Cas immunity in prokaryotes. Nature 2015, 526, 55. [Google Scholar] [CrossRef] [PubMed]
  135. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Shen, B.; Zhang, W.; Zhang, J.; Zhou, J.; Wang, J.; Chen, L.; Wang, L.; Hodgkins, A.; Iyer, V.; Huang, X.; et al. Efficient genome modification by CRISPR-Cas9 nickase with minimal off-target effects. Nat. Methods 2014, 11, 399–402. [Google Scholar] [CrossRef] [PubMed]
  137. Chu, V.T.; Weber, T.; Wefers, B.; Wurst, W.; Sander, S.; Rajewsky, K.; Kühn, R. Increasing the efficiency of homology-directed repair for CRISPR-Cas9-induced precise gene editing in mammalian cells. Nat. Biotechnol. 2015, 33, 543–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. de Santa-Inez, D.C.; Fuziwara, C.S.; Saito, K.C.; Kimura, E.T. Targeting the Highly Expressed microRNA miR-146b with CRISPR/Cas9n Gene Editing System in Thyroid Cancer. Int. J. Mol. Sci. 2021, 22, 7992. [Google Scholar] [CrossRef]
  139. Jiang, F.-N.; Liang, Y.-X.; Wei, W.; Zou, C.-Y.; Chen, G.-X.; Wan, Y.-P.; Liu, Z.-Z.; Yang, Y.; Han, Z.-D.; Zhu, J.-G.; et al. Functional classification of prostate cancer-associated miRNAs through CRISPR/Cas9-mediated gene knockout. Mol. Med. Rep. 2020, 22, 3777–3784. [Google Scholar] [CrossRef]
  140. Chang, H.; Yi, B.; Ma, R.; Zhang, X.; Zhao, H.; Xi, Y. CRISPR/cas9, a novel genomic tool to knock down microRNA in vitro and in vivo. Sci. Rep. 2016, 6, 22312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Yoshino, H.; Yonemori, M.; Miyamoto, K.; Tatarano, S.; Kofuji, S.; Nohata, N.; Nakagawa, M.; Enokida, H. microRNA-210-3p depletion by CRISPR/Cas9 promoted tumorigenesis through revival of TWIST1 in renal cell carcinoma. Oncotarget 2017, 8, 20881–20894. [Google Scholar] [CrossRef]
  142. Choudhury, S.R.; Cui, Y.; Lubecka, K.; Stefanska, B.; Irudayaraj, J. CRISPR-dCas9 mediated TET1 targeting for selective DNA demethylation at BRCA1 promoter. Oncotarget 2016, 7, 46545. [Google Scholar] [CrossRef] [Green Version]
  143. Okada, M.; Kanamori, M.; Someya, K.; Nakatsukasa, H.; Yoshimura, A. Stabilization of Foxp3 expression by CRISPR-dCas9-based epigenome editing in mouse primary T cells. Epigenetics Chromatin 2017, 10, 24. [Google Scholar] [CrossRef]
  144. Wang, H.; Guo, R.; Du, Z.; Bai, L.; Li, L.; Cui, J.; Li, W.; Hoffman, A.R.; Hu, J.-F. Epigenetic targeting of Granulin in hepatoma cells by synthetic CRISPR dCas9 epi-suppressors. Mol. Ther. Nucleic Acids 2018, 11, 23–33. [Google Scholar] [CrossRef] [Green Version]
  145. Perez-Pinera, P.; Kocak, D.D.; Vockley, C.M.; Adler, A.F.; Kabadi, A.M.; Polstein, L.R.; Thakore, P.I.; Glass, K.A.; Ousterout, D.G.; Leong, K.W.; et al. RNA-guided gene activation by CRISPR-Cas9–based transcription factors. Nat. Methods 2013, 10, 973–976. [Google Scholar] [CrossRef] [Green Version]
  146. Polstein, L.R.; Gersbach, C.A. A light-inducible CRISPR-Cas9 system for control of endogenous gene activation. Nat. Chem. Biol. 2015, 11, 198–200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Hilton, I.B.; D’Ippolito, A.M.; Vockley, C.M.; Thakore, P.I.; Crawford, G.E.; Reddy, T.E.; Gersbach, C.A. Epigenome editing by a CRISPR-Cas9-based acetyltransferase activates genes from promoters and enhancers. Nat. Biotechnol. 2015, 33, 510–517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Thakore, P.I.; Black, J.B.; Hilton, I.B.; Gersbach, C.A. Editing the epigenome: Technologies for programmable transcription and epigenetic modulation. Nat. Methods 2016, 13, 127–137. [Google Scholar] [CrossRef] [PubMed]
  149. Chakraborty, S.; Ji, H.; Kabadi, A.M.; Gersbach, C.A.; Christoforou, N.; Leong, K.W. A CRISPR/Cas9-Based System for Reprogramming Cell Lineage Specification. Stem Cell Rep. 2014, 3, 940–947. [Google Scholar] [CrossRef] [Green Version]
  150. Thakore, P.I.; D’Ippolito, A.M.; Song, L.; Safi, A.; Shivakumar, N.K.; Kabadi, A.M.; Reddy, T.E.; Crawford, G.E.; Gersbach, C.A. Highly specific epigenome editing by CRISPR-Cas9 repressors for silencing of distal regulatory elements. Nat. Methods 2015, 12, 1143–1149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Maddalo, D.; Manchado, E.; Concepcion, C.P.; Bonetti, C.; Vidigal, J.A.; Han, Y.-C.; Ogrodowski, P.; Crippa, A.; Rekhtman, N.; de Stanchina, E.; et al. In vivo engineering of oncogenic chromosomal rearrangements with the CRISPR/Cas9 system. Nature 2014, 516, 423–427. [Google Scholar] [CrossRef] [Green Version]
  152. Sakuma, T.; Nishikawa, A.; Kume, S.; Chayama, K.; Yamamoto, T. Multiplex genome engineering in human cells using all-in-one CRISPR/Cas9 vector system. Sci. Rep. 2014, 4, 5400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Bordin, M.; D’Atri, F.; Guillemot, L.; Citi, S. Histone Deacetylase Inhibitors Up-Regulate the Expression of Tight Junction Proteins11Swiss Cancer League, Swiss National Science Foundation, Ministry for Italian University and Research, ERASMUS Program (M. Bordin), and Roche Research Foundation fellowship (L. Guillemot). Mol. Cancer Res. 2004, 2, 692–701. [Google Scholar] [PubMed]
  154. Yang, S.; Zhang, J.; Zhang, Y.; Wan, X.; Zhang, C.; Huang, X.; Huang, W.; Pu, H.; Pei, C.; Wu, H. KDM1A triggers androgen-induced miRNA transcription via H3K4me2 demethylation and DNA oxidation. Prostate 2015, 75, 936–946. [Google Scholar] [CrossRef]
  155. Georgakilas, G.; Perdikopanis, N.; Hatzigeorgiou, A.G. Identifying Pri-miRNA Transcription Start Sites. In miRNA Biogenesis; Springer: Berlin/Heidelberg, Germany, 2018; pp. 11–31. [Google Scholar]
  156. Zhao, Y.; Dai, Z.; Liang, Y.; Yin, M.; Ma, K.; He, M.; Ouyang, H.; Teng, C.-B. Sequence-specific inhibition of microRNA via CRISPR/CRISPRi system. Sci. Rep. 2014, 4, 3943. [Google Scholar] [CrossRef] [Green Version]
  157. Gilbert, L.A.; Larson, M.H.; Morsut, L.; Liu, Z.; Brar, G.A.; Torres, S.E.; Stern-Ginossar, N.; Brandman, O.; Whitehead, E.H.; Doudna, J.A.; et al. CRISPR-mediated modular RNA-guided regulation of transcription in eukaryotes. Cell 2013, 154, 442–451. [Google Scholar] [CrossRef] [Green Version]
  158. Zetsche, B.; Gootenberg, J.S.; Abudayyeh, O.O.; Slaymaker, I.M.; Makarova, K.S.; Essletzbichler, P.; Volz, S.E.; Joung, J.; van der Oost, J.; Regev, A.; et al. Cpf1 Is a Single RNA-Guided Endonuclease of a Class 2 CRISPR-Cas System. Cell 2015, 163, 759–771. [Google Scholar] [CrossRef] [Green Version]
  159. Ran, F.A.; Hsu, P.D.; Lin, C.-Y.; Gootenberg, J.S.; Konermann, S.; Trevino, A.E.; Scott, D.A.; Inoue, A.; Matoba, S.; Zhang, Y.; et al. Double Nicking by RNA-Guided CRISPR Cas9 for Enhanced Genome Editing Specificity. Cell 2013, 154, 1380–1389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Maresca, M.; Lin, V.G.; Guo, N.; Yang, Y. Obligate ligation-gated recombination (ObLiGaRe): Custom-designed nuclease-mediated targeted integration through nonhomologous end joining. Genome Res. 2013, 23, 539–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Zhang, X.; Wang, W.; Shan, L.; Han, L.; Ma, S.; Zhang, Y.; Hao, B.; Lin, Y.; Rong, Z. Gene activation in human cells using CRISPR/Cpf1-p300 and CRISPR/Cpf1-SunTag systems. Protein Cell 2018, 9, 380–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Liu, Y.; Han, J.; Chen, Z.; Wu, H.; Dong, H.; Nie, G. Engineering cell signaling using tunable CRISPR–Cpf1-based transcription factors. Nat. Commun. 2017, 8, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Ozsolak, F.; Poling, L.L.; Wang, Z.; Liu, H.; Liu, X.S.; Roeder, R.G.; Zhang, X.; Song, J.S.; Fisher, D.E. Chromatin structure analyses identify miRNA promoters. Genes Dev. 2008, 22, 3172–3183. [Google Scholar] [CrossRef] [Green Version]
  164. Strecker, J.; Jones, S.; Koopal, B.; Schmid-Burgk, J.; Zetsche, B.; Gao, L.; Makarova, K.S.; Koonin, E.V.; Zhang, F. Engineering of CRISPR-Cas12b for human genome editing. Nat. Commun. 2019, 10, 1–8. [Google Scholar] [CrossRef] [PubMed]
  165. Engreitz, J.; Abudayyeh, O.; Gootenberg, J.; Zhang, F. CRISPR Tools for Systematic Studies of RNA Regulation. Cold Spring Harb. Perspect. Biol. 2019, 11, a035386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Cox, D.B.T.; Gootenberg, J.S.; Abudayyeh, O.O.; Franklin, B.; Kellner, M.J.; Joung, J.; Zhang, F. RNA editing with CRISPR-Cas13. Science 2017, 358, 1019–1027. [Google Scholar] [CrossRef] [Green Version]
  167. Abudayyeh, O.O.; Gootenberg, J.S.; Essletzbichler, P.; Han, S.; Joung, J.; Belanto, J.J.; Verdine, V.; Cox, D.B.T.; Kellner, M.J.; Regev, A.; et al. RNA targeting with CRISPR–Cas13. Nature 2017, 550, 280–284. [Google Scholar] [CrossRef] [Green Version]
  168. Brennecke, J.; Stark, A.; Russell, R.B.; Cohen, S.M. Principles of MicroRNA–Target Recognition. PLoS Biol. 2005, 3, e85. [Google Scholar] [CrossRef]
  169. O’Donnell, K.A.; Wentzel, E.A.; Zeller, K.I.; Dang, C.V.; Mendell, J.T. c-Myc-regulated microRNAs modulate E2F1 expression. Nature 2005, 435, 839–843. [Google Scholar] [CrossRef]
  170. Olena, A.F.; Patton, J.G. Genomic organization of microRNAs. J. Cell. Physiol. 2010, 222, 540–545. [Google Scholar] [CrossRef] [Green Version]
  171. Ratnadiwakara, M.; Mohenska, M.; Änkö, M.-L. Splicing factors as regulators of miRNA biogenesis–links to human disease. In Seminars in Cell & Developmental Biology; Elsevier: Amsterdam, The Netherlands, 2018; Volume 79, pp. 113–122. [Google Scholar]
  172. Zhang, X.-H.; Tee, L.Y.; Wang, X.-G.; Huang, Q.-S.; Yang, S.-H. Off-target Effects in CRISPR/Cas9-mediated Genome Engineering. Mol. Ther. Nucleic Acids 2015, 4, e264. [Google Scholar] [CrossRef] [PubMed]
  173. Kocher, T.; Peking, P.; Klausegger, A.; Murauer, E.M.; Hofbauer, J.P.; Wally, V.; Lettner, T.; Hainzl, S.; Ablinger, M.; Bauer, J.W.; et al. Cut and Paste: Efficient Homology-Directed Repair of a Dominant Negative KRT14 Mutation via CRISPR/Cas9 Nickases. Mol. Ther. 2017, 25, 2585–2598. [Google Scholar] [CrossRef] [Green Version]
  174. Zlotorynski, E. Epigenetics: Characterizing enhancers with dCas9. Nat. Rev. Mol. Cell Biol. 2015, 16, 266. [Google Scholar] [CrossRef] [PubMed]
  175. Gong, J.; Tang, D.; Leong, K.W. CRISPR/dCas9-mediated cell differentiation. Curr. Opin. Biomed. Eng. 2018, 7, 9–15. [Google Scholar] [CrossRef]
  176. Feng, J.; Liu, T.; Qin, B.; Zhang, Y.; Liu, X.S. Identifying ChIP-seq enrichment using MACS. Nat. Protoc. 2012, 7, 1728–1740. [Google Scholar] [CrossRef] [Green Version]
  177. Liu, X.S.; Wu, H.; Ji, X.; Stelzer, Y.; Wu, X.; Czauderna, S.; Shu, J.; Dadon, D.; Young, R.A.; Jaenisch, R. Editing DNA methylation in the mammalian genome. Cell 2016, 167, 233–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Xu, X.; Tao, Y.; Gao, X.; Zhang, L.; Li, X.; Zou, W.; Ruan, K.; Wang, F.; Xu, G.; Hu, R. A CRISPR-based approach for targeted DNA demethylation. Cell Discov. 2016, 2, 16009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Li, Y.; Tollefsbol, T.O. DNA methylation detection: Bisulfite genomic sequencing analysis. In Epigenetics Protocols; Springer: Berlin/Heidelberg, Germany, 2011; pp. 11–21. [Google Scholar]
  180. Jiang, Q.; Meng, X.; Meng, L.; Chang, N.; Xiong, J.; Cao, H.; Liang, Z. Small indels induced by CRISPR/Cas9 in the 5′ region of microRNA lead to its depletion and Drosha processing retardance. RNA Biol. 2014, 11, 1243–1249. [Google Scholar] [CrossRef] [Green Version]
  181. Huo, W.; Zhao, G.; Yin, J.; Ouyang, X.; Wang, Y.; Yang, C.; Wang, B.; Dong, P.; Wang, Z.; Watari, H. Lentiviral CRISPR/Cas9 vector mediated miR-21 gene editing inhibits the epithelial to mesenchymal transition in ovarian cancer cells. J. Cancer 2017, 8, 57. [Google Scholar] [CrossRef]
  182. Abdollah, N.A.; Kumitaa, T.D.; Narazah, M.Y.; Razak, S.R.A. Sequence-specific inhibition of microRNA-130a gene by CRISPR/Cas9 system in breast cancer cell line. J. Phys. Conf. Ser. 2017, 851, 012037. [Google Scholar] [CrossRef] [Green Version]
  183. Li, X.; Chen, W.; Zeng, W.; Wan, C.; Duan, S.; Jiang, S. microRNA-137 promotes apoptosis in ovarian cancer cells via the regulation of XIAP. Br. J. Cancer 2017, 116, 66–76. [Google Scholar] [CrossRef] [Green Version]
  184. Chen, W.; Li, Q.; Du, J.; Li, X.; Jiang, S.; He, Y. Establishment of an miR-137-knockout cell model using CRISPR/Cas9 genome editing. Oncol. Lett. 2018, 16, 4027–4032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Nayak, S.; Aich, M.; Kumar, A.; Sengupta, S.; Bajad, P.; Dhapola, P.; Paul, D.; Narta, K.; Purkrait, S.; Mehani, B.; et al. Novel internal regulators and candidate miRNAs within miR-379/miR-656 miRNA cluster can alter cellular phenotype of human glioblastoma. Sci. Rep. 2018, 8, 7673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Hannafon, B.N.; Cai, A.; Calloway, C.L.; Xu, Y.-F.; Zhang, R.; Fung, K.-M.; Ding, W.-Q. miR-23b and miR-27b are oncogenic microRNAs in breast cancer: Evidence from a CRISPR/Cas9 deletion study. BMC Cancer 2019, 19, 642. [Google Scholar] [CrossRef]
  187. Akdemir, B.; Nakajima, Y.; Inazawa, J.; Inoue, J. miR-432 Induces NRF2 Stabilization by Directly Targeting KEAP1. Mol. Cancer Res. 2017, 15, 1570–1578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Fuziwara, C.S.; Saito, K.C.; Kimura, E.T. Thyroid Follicular Cell Loss of Differentiation Induced by MicroRNA miR-17-92 Cluster Is Attenuated by CRISPR/Cas9n Gene Silencing in Anaplastic Thyroid Cancer. Thyroid 2019, 30, 81–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Fang, L.-L.; Wang, X.-H.; Sun, B.-F.; Zhang, X.-D.; Zhu, X.-H.; Yu, Z.-J.; Luo, H. Expression, regulation and mechanism of action of the miR-17-92 cluster in tumor cells. Int. J. Mol. Med. 2017, 40, 1624–1630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Pajic, M.; Froio, D.; Daly, S.; Doculara, L.; Millar, E.; Graham, P.H.; Drury, A.; Steinmann, A.; Bock, C.E.; de Boulghourjian, A.; et al. miR-139-5p Modulates Radiotherapy Resistance in Breast Cancer by Repressing Multiple Gene Networks of DNA Repair and ROS Defense. Cancer Res. 2018, 78, 501–515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Dragomir, M.P.; Knutsen, E.; Calin, G.A. Classical and noncanonical functions of miRNAs in cancers. Trends Genet. 2021. [Google Scholar] [CrossRef] [PubMed]
Figure 1. miRNA loci amplification resulted in repeating miRNA sequences thus increasing basal expression of miRNA.
Figure 1. miRNA loci amplification resulted in repeating miRNA sequences thus increasing basal expression of miRNA.
Biomedicines 10 00915 g001
Figure 2. Single base disposition at miRNA locus.
Figure 2. Single base disposition at miRNA locus.
Biomedicines 10 00915 g002
Figure 3. (A): CpG site enrichment on miRNA locus resulting in miRNA suppression. CpG methylation brought about by DNMTs methylates DNA sequence in proximity to miRNA locus or direct locus methylation leads to suppression of miRNA expression. H3C: methyl group; (B): Histone state on miRNA locus controls accessibility of TFs. Histone modification leads to the formation of open or closed chromatin structure which in turn affect TFs binding onto miRNA locus where this would result in either miRNA activation or suppression depending on the transcriptional activity of the bounded factors. H3C: methyl group.
Figure 3. (A): CpG site enrichment on miRNA locus resulting in miRNA suppression. CpG methylation brought about by DNMTs methylates DNA sequence in proximity to miRNA locus or direct locus methylation leads to suppression of miRNA expression. H3C: methyl group; (B): Histone state on miRNA locus controls accessibility of TFs. Histone modification leads to the formation of open or closed chromatin structure which in turn affect TFs binding onto miRNA locus where this would result in either miRNA activation or suppression depending on the transcriptional activity of the bounded factors. H3C: methyl group.
Biomedicines 10 00915 g003
Figure 4. Tools for regulating miRNA; miRNA sponges use overexpression plasmids to produce miRNA target sequence which sequesters miRNAs reducing binding on target mRNA thus increasing target mRNA expression whereas antagomir is a complementary sequence of short RNA that binds mRNA via base pairing, the use of antagomir prevents miRNA from inducing mRNA cleavage thus resulting in increase in target expression.
Figure 4. Tools for regulating miRNA; miRNA sponges use overexpression plasmids to produce miRNA target sequence which sequesters miRNAs reducing binding on target mRNA thus increasing target mRNA expression whereas antagomir is a complementary sequence of short RNA that binds mRNA via base pairing, the use of antagomir prevents miRNA from inducing mRNA cleavage thus resulting in increase in target expression.
Biomedicines 10 00915 g004
Figure 5. miRNA mimic uses short RNA sequence which mimics target miRNA bases sequence, this acts as a tool in inducing miRNA activity which would result in downregulation of target mRNA.
Figure 5. miRNA mimic uses short RNA sequence which mimics target miRNA bases sequence, this acts as a tool in inducing miRNA activity which would result in downregulation of target mRNA.
Biomedicines 10 00915 g005
Figure 6. (A) Using dCas9 to block transcription elongation. TFs: transcription factors; Pol II: RNA polymerase 2; (B) Using the nickase strategy to cut amplified loci. HDR: homology-directed repair; (C) Relieving the epigenetics state of miRNA locus to increase gene transcription. Act: Acetyl group; p300: Histone modifier; (D) Modulation of the CpG methylation to induce miRNA expression. Modulation of the CpG methylation to induce miRNA expression. H3C: methyl group.
Figure 6. (A) Using dCas9 to block transcription elongation. TFs: transcription factors; Pol II: RNA polymerase 2; (B) Using the nickase strategy to cut amplified loci. HDR: homology-directed repair; (C) Relieving the epigenetics state of miRNA locus to increase gene transcription. Act: Acetyl group; p300: Histone modifier; (D) Modulation of the CpG methylation to induce miRNA expression. Modulation of the CpG methylation to induce miRNA expression. H3C: methyl group.
Biomedicines 10 00915 g006aBiomedicines 10 00915 g006b
Figure 7. General workflow for studying miRNA dysfunction in cancer. Depending on experimental design miRNA study could be assessed via knock-in or loss of function study. Several CRISPR-based methods are available as listed in Figure for the desired experiments.
Figure 7. General workflow for studying miRNA dysfunction in cancer. Depending on experimental design miRNA study could be assessed via knock-in or loss of function study. Several CRISPR-based methods are available as listed in Figure for the desired experiments.
Biomedicines 10 00915 g007
Table 1. Table showing various mechanisms of dysregulation of miRNA in cancers.
Table 1. Table showing various mechanisms of dysregulation of miRNA in cancers.
miRNAMechanism of DysregulationConsequenceReferences
miR-650Loci amplificationInverse correlation was observed between miR-650 and tumour supressor genes ING4 and NDRG2[77]
miR-21Loci amplificationmiR-21 overexpression leads to PTEN suppression[78]
miR-4288DeletionLoss of miRNA in prostate cancer, miRNA directly represses metastatic/invasion genes MMP16 and ROCK1[79]
miR-3613DeletionmiR-3613 was found to be lower in breast cancer. Gain of function reveals miR-3613 to regulate PAFAH1B2 and PDK3 blocking oncogenesis[80]
miR-379Base substitutionHotspot mutation commonly occurring in lung adenocarcinoma[81]
miR-142-3pEpigenetic suppression via DNMT recruitmentHypermethylation of miR-142 leads to unfavorable prognosis in nasopharyngeal carcinoma [82]
miR-338-5p/421EZH2 mediated suppression via DNA methylationPresence of CpG marks on primary prostate cancer. Ectopic expression reveals suppression in prostate cancer growth [83]
miR-17-92/106b-25CMYC drivenCMYC drive the expression of these miRNA clusters, inhibition of cMYC activators resulted in suppression of these clusters in hepatocellular carcinoma[84]
miR-122CMYC drivenCMYC oncogene overexpression in hepatocellular carcinoma activates miR-122 via direct promoter binding driving oncogenesis[85]
miR-455-3pReside in host gene driven by p53miRNA involves in cancer quiescence via p53 mediation[86]
Table 2. Various conventional techniques to study miRNA function.
Table 2. Various conventional techniques to study miRNA function.
Method of StudymiRNAMechanismOutcomeReferences
miRNA spongesmiR-21Synthetic RNA sponge was designed to bear sequence complementary to miR-21Downregulated proteins due to miR-21 overexpression were restored[124]
miR-21 and miR-93Single RNA sponges bearing multiple complementary sites against target miRNAsTargetting oncomiRs effectively induce apoptosis and blocked proliferation of esophageal carcinoma[125]
miR-223RNA sponge Expressing DNA plasmid was used. RNA circularization was imposed using slicing acceptor and donor siteSponges effectively sequester endogenous miRNA in T-ALL cells effective restoration of miR target genes[126]
Antagomirs/Antisense Oligo nucleotidemiR-155-5pTransfection of breast cancer cell lines with antagonist against miRDownregulation of miR leads to increase in breast cancer sensitivity against cetuximab[127]
miR-125a-5pTransfection of gastric cancer cells with antagomir Restoration of miR suppressed genes was observed, suppression leads to the suppression in EMT of gastric cancer[128]
miR-155Delivery of antagomir into MCF-7 via attachment to gold nano particleElevation in miR target gene T53INP1 was observed stimulating apoptosis of breast cancer cells[129]
miRNA mimicsmiR-27aOverexpression of miR via mimics transfection
Mimics used are as standard siRNA sizes
Overexpression of miR-27a via mimic alleviates cancer characteristics and sensitizes breast cancer towards anticancer drugs [130]
miR-150miRNA sequence was expressed using pEZX-MR via lentiviral deliveryMimic expression induces apoptosis in multiple leukemic cell lines[131]
Table 3. Table showing miRNA target using CRISPR/Cas9 via different modes at miRNA processing stage.
Table 3. Table showing miRNA target using CRISPR/Cas9 via different modes at miRNA processing stage.
microRNA TargetTarget SiteCRISPR System, DeliveryModelOutcome
miR-93 [180]5′ Drosha Processing siteCRISPR/Cas9, LipofectionHuman Cervical Cancer (HeLa)Almost no detection of mature miR-93 Accumulation of primary miR-93 transcript suggest impairment in Drosha processing
miR-21 [181]20nt sequence adjacent to PAM (NGG)CRISPR/Cas9, Lentiviral VectorHuman Ovarian Adenocarcinoma (SKOV3 and OVCR3)Significant reduction in mature miR-21 expression was observed
miR-130a [182]5p and 3p Seed Sequence
Stem Loop (Dicer binding Site)
CRISPR/Cas9, LipofectionHuman Breast Cancer (MCF7)Significant reduction was observed when using Cas9 targeting from the 5p region No significant difference in miR-130a expression was observed when targeting either the 3p or the Stem Loop sequence
miR-137 [183,184]Nucleotide sequence upstream of 5′ PAM (NGG)CRISPR/Cas9, Lentiviral VectorHuman Ovarian Carcinoma (A2780)Significant reduction in mature miR-137 expression was observed. Deletion and insertion mutation detected from single-cell expanded colonies.
miR-379/miR-656 cluster [185]dCas9 fused to VP-64 docking on the miRNA locus for induction of miRNA gene expression. CRISPR/dCas9, LipofectionHuman GlioblastomaIncrease in expression of miRNA within the miR-379/miR-656 cluster post-dCas9-VP64 gene induction.
miR-23b and miR-27b [186]Annotated Stem-loop regionCas9/Lentiviral TransductionHuman Breast Cancer
(MCF7)
Significant reduction of miR-23b and miR-27b transcripts was observed
miR-423 [187]miR-423 locus Cas9/LipofectionHuman Cervical Cancer (HeLa)Significant knockdown of miR-423 transcripts was observed
miR-17-92 [188]miR-17-92 5p loopCRISPR/Cas9 nickasesAnaplastic Thyroid CancerKnockdown of clusters was observed
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Azlan, A.; Rajasegaran, Y.; Kang Zi, K.; Rosli, A.A.; Yik, M.Y.; Yusoff, N.M.; Heidenreich, O.; Moses, E.J. Elucidating miRNA Function in Cancer Biology via the Molecular Genetics’ Toolbox. Biomedicines 2022, 10, 915. https://0-doi-org.brum.beds.ac.uk/10.3390/biomedicines10040915

AMA Style

Azlan A, Rajasegaran Y, Kang Zi K, Rosli AA, Yik MY, Yusoff NM, Heidenreich O, Moses EJ. Elucidating miRNA Function in Cancer Biology via the Molecular Genetics’ Toolbox. Biomedicines. 2022; 10(4):915. https://0-doi-org.brum.beds.ac.uk/10.3390/biomedicines10040915

Chicago/Turabian Style

Azlan, Adam, Yaashini Rajasegaran, Khor Kang Zi, Aliaa Arina Rosli, Mot Yee Yik, Narazah Mohd Yusoff, Olaf Heidenreich, and Emmanuel Jairaj Moses. 2022. "Elucidating miRNA Function in Cancer Biology via the Molecular Genetics’ Toolbox" Biomedicines 10, no. 4: 915. https://0-doi-org.brum.beds.ac.uk/10.3390/biomedicines10040915

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

Article Metrics

Back to TopTop