Skip to main content

Targeting miRNA by CRISPR/Cas in cancer: advantages and challenges

Abstract

Clustered regulatory interspaced short palindromic repeats (CRISPR) has changed biomedical research and provided entirely new models to analyze every aspect of biomedical sciences during the last decade. In the study of cancer, the CRISPR/CRISPR-associated protein (Cas) system opens new avenues into issues that were once unknown in our knowledge of the noncoding genome, tumor heterogeneity, and precision medicines. CRISPR/Cas-based gene-editing technology now allows for the precise and permanent targeting of mutations and provides an opportunity to target small non-coding RNAs such as microRNAs (miRNAs). However, the development of effective and safe cancer gene editing therapy is highly dependent on proper design to be innocuous to normal cells and prevent introducing other abnormalities. This study aims to highlight the cutting-edge approaches in cancer-gene editing therapy based on the CRISPR/Cas technology to target miRNAs in cancer therapy. Furthermore, we highlight the potential challenges in CRISPR/Cas-mediated miRNA gene editing and offer advanced strategies to overcome them.

Background

In 1987, the first instance of clustered regulatory interspaced short palindromic repeats (CRISPR) was found in the bacteria Escherichia coli K12 [1]. For the past 20 years, these palindromic repeats have been discovered in approximately 40% of bacteria and 90% of archaea [2]. CRISPR has repeat sequences that are spaced by exogenous nucleotides from plasmids or viruses that have invaded, and its loci are frequently surrounded by some related endonucleases, like CRISPR-associated protein (Cas). First, precursor CRISPR RNAs (pre-crRNAs) are produced from CRISPR. The resulting crRNAs bind to the Cas protein to form a complex that can activate the transcription of certain DNA regions [3]. Although the main activity of this ribonucleoprotein (RNP) complex is to cleave specific DNA locus, specified by the crRNA sequence, with the nuclease activity of the Cas protein. There are three stages to the immune response in all known CRISPR/Cas systems: 1) CRISPR arrays can undergo adaptation and spacer acquisition, in which a fragment of the invading genome is added to the existing gene, 2) mature crRNAs [guide RNA (gRNAs)] are expressed as a result of the CRISPR array processing, and 3) interference, wherein the gRNAs direct Cas proteins to the target location of the invaded genome for destruction or cleavage [4, 5].

There are various classes and types of CRISPR systems, the widest one is class 2, CRISPR/Cas9. Here, in CRISPR/Cas9, the Cas protein works in conjunction with a chimeric single-guide RNA (sgRNA) made from crRNA and tracrRNA. TracrRNA is necessary for Cas nuclease activity, while crRNA detects and binds sequences next to the protospacer adjacent motif (PAM), 5′-NGG-3′, on the target DNA sequences [6]. The target DNA sequence is complementary to the first 20 nucleotides of the sgRNA, which are then followed by a sequence known as PAM, which is generally NGG [7].

The CRISPR/Cas system has potential applications in medicine, including diagnostics, therapeutics, and drug screening. Despite the growing popularity of the CRISPR/Cas technology for gene editing, being used in studying microRNAs (miRNAs) remains mostly undefined [8].

In addition, studies revealed that applying CRISPR/Cas is significantly less expensive, has a lower chance of contamination, and is more accruable and specific in its ability to target miRNAs in cancer therapy, when compared with the current miRNA studying approaches [9].

Small non-coding RNAs known as miRNAs influence gene expression by acting as either transcriptional regulators or translational repressors of their downstream target genes [10]. In mammals, it is expected that almost half of all protein-coding genes’ activity is regulated by miRNAs, which are highly conserved non-coding regulatory factors [11]. In human malignancies, miRNA expression is dysregulated by a number of processes, including miRNA gene amplification or deletion, improper miRNA transcriptional regulation, dysregulated epigenetic alterations, and errors in the miRNA biogenesis machinery [12,13,14,15,16]. Furthermore, miRNAs dysregulations have been demonstrated to influence the characteristics of cancer, such as maintaining proliferative signaling, avoiding growth suppressors, apoptosis resistance [17,18,19], inducing invasion and metastasis [20], drug resistance [21], and inducing angiogenesis [22]. Therefore, there is a lot of potential for using miRNAs as diagnostic and therapeutic targets in cancer therapy. This study explores the novel insights that have been achieved due to the development of CRISPR/Cas systems as a strategy to target miRNAs in cancer therapy. Besides, we discussed the potential challenges and advanced strategies that can be applied to overcome these challenges.

Biogenesis of miRNAs and regulatory mechanisms and their role in cancer

Single-stranded, non-coding RNAs called miRNAs are derived from primary miRNA (pri-miRNA), an early transcript produced by RNA polymerase II (Pol II) [23]. Approximately 50% of the known miRNAs are made from the introns and a few exons of protein-coding genes. The other 50%, intergenic, are made from their promoters and do not depend on host genes for transcription and expression [24, 25]. Pri-miRNAs are synthesized using the same transcription steps, capping, 3′ polyadenylation, and splicing that are used to make mRNA. DNA-dependent RNA Pol II is the enzyme responsible for transcribing miRNA genes (Fig. 1). The miRNAs can sometimes be transcribed as a cluster, which is a single long transcript. Clusters can have similar or the same seed regions, which means they are a family [26, 27]. Following transcription, the nucleus produces pri-miRNAs with a characteristic stem-loop structure. The miRNA/miRNA duplex is then released after pri-miRNAs undergo two constitutive cuts, which result in pre-miRNAs. Drosha cuts pri-miRNA into pre-miRNA in the nucleus, which exportin-5 transfers to the cytoplasm [28]. Dicer uses pre-miRNA as a template to make mature, functional double-stranded (ds) miRNA [29]. After maturation, miRNA usually binds to a 3′ UTR and either destroys or suppresses mRNA translation. It has been demonstrated that a single miRNA can regulate the expression of numerous mRNAs, and each mRNA could also be controlled by different miRNAs. Based on experimental findings, in most cases, miRNAs bind to specific sequences at the 3′ UTR of their target mRNAs to trigger translational inhibition and the decapping and deadenylation of the mRNAs they target [30, 31]. It has been discovered that miRNA binding sites can also be located in other mRNA regions, such as the 5′ UTR, coding sequence, and promoter regions [32]. In particular, miRNAs inhibit gene expression when they bind to the 5′ UTR and coding sequences [33]. However, miRNAs that bind to promoter regions enhance the transcription process [34]. Nevertheless, additional study is necessary to comprehend the practical importance of this kind of interaction.

Fig. 1
figure 1

A graphical illustration of miRNA biogenesis. The pre-miRNAs have one or more incomplete hairpin structures that have a stem of about 33 base pairs. Ribonucleases Drosha and Dicer process the pri-miRNA precursor in two separate processes. In the nucleus, Drosha first cuts the pri-miRNA into a pre-miRNA about 70 nucleotides in length, which is then transferred to the cytoplasm by XPO5. The mature, functional, ds miRNA is then processed by Dicer using the pre-miRNA as a template. After maturation, the miRNA is covalently linked to RISC, a multiprotein complex that contains the AGO protein and is essential for RNA silencing. Exon 1 and exon 2 are connected together when the RNA splicing process takes place and leads to the formation of the lariat RNA (circular molecules with a short tail). Following RNA splicing and additional processing, the intron-containing spliced lariat may function as a pri-miRNA for intronic miRNA synthesis. XPO5 exportin-5, ds double-stranded, RISC RNA-induced silencing complex, AGO argonaute, ADAR adenosine deaminase RNA specific, TRBP tar RNA-binding protein, EGFR epidermal growth factor receptor

Furthermore, miRNA plays a crucial role in promoting or inhibiting cancer progression through oncogenic or tumor suppressor miRNAs (Fig. 2). Indeed, normal tissues and tumor tissues have different expressions of some types of miRNAs. Thus, from a therapeutic perspective, targeting miRNA has been considered an effective approach in cancer therapy, especially by designing specific miRNA inhibitors to target oncogenic miRNAs that are overexpressed in tumour cells [35]. Pathogenic miRNA alterations can be regulated with miRNA mimics or antagomiRs (anti-miRs), leading to adjusting the gene regulatory network and normalizing the signaling pathways in cancer cells [36].

Fig. 2
figure 2

A graphical illustration of how oncogenic and tumor suppressor miRNAs are regulated during tumorigenic events. a When oncogenic miRNAs are expressed at higher levels in malignant cells, tumor suppressor gene expression is lowered either as a result of mRNA degradation or hypermethylation. b Oncogenic miRNA expression may be increased by decreasing the expression level of tumor suppressor miRNAs. Both oncogenic and tumor suppressor miRNAs contribute to tumorigenesis by promoting a variety of malignant phenotypes, including cell development, anti-apoptotic activity, invading, angiogenic, and spreading. RISC RNA-induced silencing complex, ORF open reading frame

When miRNA binds to a 3′ UTR, it either destroys the mRNA or inhibits its translation. The degree of miRNA complementarity to the 3′ UTR determines the level of mRNA degradation or translational repression [37].

MiRNAs are a type of non-coding RNA that influence gene expression after transcription has already taken place by interacting with the 3′ UTR of mRNA. Friedman et al. [38] analyzed more than 45,000 miRNA target locations in the human 3′ UTR region, and they found that miRNA regulates roughly 60% of human protein-coding genes.

Dysregulation of miRNA expression has been associated with various pathological conditions, including cancer [39,40,41]. In cancer, miRNAs can act as either tumor suppressors or oncogenes, depending on the target genes they regulate. The role of miRNAs in cancer progression is complex and involves multiple mechanisms. Some miRNAs have crucial roles in cancer progression, including the control of oncogenes and tumor suppressor genes. For instance, miR-21 is an oncogenic miRNA that targets tumor suppressor genes such as PTEN and PDCD4 [42], while miR-34a is a tumor suppressor miRNA that targets oncogenes such as MYC and BCL2 [43]. Further, miRNAs can promote cancer cell proliferation and survival by targeting genes involved in cell cycle regulation, cell death, and DNA damage response. For example, the miR-17-92 cluster, which is up-regulated in many cancers, promotes cancer cell proliferation by targeting the tumor suppressor gene PTEN [44]. Moreover, several miRNAs have been shown to regulate epithelial-to-mesenchymal transition (EMT) by targeting genes involved in cell–cell adhesion and cytoskeleton organization. For example, miR-200 family miRNAs inhibit EMT by targeting the transcription factors zinc finger E-box-binding homeobox 1 (ZEB1) and ZEB2 [45]. Likewise, miRNAs have been found to regulate angiogenesis by targeting genes involved in angiogenic signaling pathways. For example, miR-126 inhibits angiogenesis by targeting vascular endothelial growth factor A (VEGF-A) and phosphatidylinositol 3-kinase regulatory subunit 2 (PIK3R2) [46]. Additionally, miRNAs can also regulate immune responses by targeting genes involved in immune signaling pathways. For example, miR-155 promotes inflammation by targeting the negative regulator of nuclear factor-κB (NF-κB), a suppressor of cytokine signaling 1 (SOCS1) [47, 48]. MiRNAs influence alternative splicing and chromatin remodeling [49]. Dysregulated epigenetic changes of miRNA may induce tumors [50].

It has been discovered that miRNAs play an important role in cancers beyond the intracellular level, for example, in extracellular fluids, either as free circulating molecules or enclosed in exosomes. These extracellular molecules play a crucial role in cell signaling, and they are capable of traveling extensive distances to exert their effects on recipient cells, particularly immune cells in the tumor microenvironment [51].

Overall, miRNAs play critical roles in cancer progression by regulating various cellular processes. Understanding the precise mechanisms underlying miRNA dysregulation in cancer is essential for developing effective miRNA-based therapies for cancer.

Therefore, therapeutically, miRNA has been perceived as a useful method in cancer therapy, particularly in the construction of specific miRNA inhibitors to target oncogenic miRNAs that are overexpressed in tumor cells [52]. Thus, correction of miRNA abnormalities in cancer cells using gene editing tools such as CRISPR/Cas, can restore normal function to the cells’ gene regulatory networks and signaling cascades [53].

Challenges of current miRNA-based cancer therapy

An innovative path for cancer research and treatment would be possible through the identification of novel therapeutic drugs that can specifically inhibit oncogenic miRNAs [54].

Since an imbalance in miRNA expression levels is associated with the development of cancer, miRNA-based cancer therapies are developed with two distinct guiding principles: the suppression of miRNAs that are overexpressed and the restoration of tumor suppressor miRNA function (Fig. 3) [55]. To restore the expression level of tumor-suppressive miRNAs and improve the function of endogenous miRNAs, miRNA mimics and other small compounds are typically used to repair miRNA function. On the other hand, overexpressed miRNAs can be inhibited using small-molecule inhibitors, antagomiRs, and miRNA sponges that have been specifically created to target particular oncogenic miRNAs that are overexpressed in cancer cells [56]. However, each of these techniques, as summarized in Table 1 [57,58,59,60,61,62,63,64,65,66,67,68,69,70,71], has its limitation, and requires further study to become more effective and less toxic.

Fig. 3
figure 3

Therapeutic use of miRNAs in cancer treatment. miRNA replacement therapies or oncogenic miRNA inhibition are the two primary current methods utilized to prevent the overexpression and functions of miRNAs. miRNA replacement therapy such as ligand conjugated miRNAs, liposomes, miRNA mimics, and viral vectors are used to suppress the oncogenic miRNAs. Furthermore, miRNA inhibition therapy includes small molecule inhibitors, miRNA sponge, antisense, and CRISPR/Cas9 which inhibit the oncogenic miRNA’s function. miRNAs microRNAs, CRISPR/Cas9 clustered regularly interspaced short palindromic repeats/CRISPR-associated protein 9

Table 1 Advantages and disadvantages of current miRNA inhibitors and enhancer

Due to the limited cellular uptake properties used in the current delivery carriers, the primary drawback of miRNA-based therapy is delivery efficiency [72]. Additionally, to accomplish its medicinal objective, it requires the proper delivery of the particles into and via the body’s complex circulatory system as well as through the cell membranes of the various tissues.

The second challenge with miRNA-based therapy is the specificity and off-target effects of miRNA. As previously mentioned, one of the future challenges of miRNA is “multi-targeting” which gives the advance to treat diseases by affecting multiple pathogenesis-correlated targets. However, since the binding between miRNA and target mRNA requires only a partial complementarity, the risk of off-target effects is very high [73].

The third challenge of current miRNA base therapy is the miRNA-induced toxicity, as a result of the off-target occurrence. If miRNAs transcriptionally control the expression of non-target genes, such as drug-metabolizing enzymes, alterations may occur and lead to toxicity. Deregulating the expression of cytochrome P450s (CYPs) and bile acid synthase cholesterol 7 alpha-hydroxylase (CYP7A1) by particular miRNA may change drug metabolism and increase drug accumulation, as in the case of CYPs and bile acid synthase CYP7A1 [74]. The approach of miRNA agomir has shown some significant drawbacks [64]. A human clinical trial demonstrates that the use of MRX34, which is a liposomal mimic of miR-34a, to treat liver cancer in humans has caused clinical toxicity in some individuals. Furthermore, the trial ended as a result of severe immune-mediated adverse impact, which resulted in the loss of four volunteers [75].

The rapid clearance of a naked nucleotide in the bloodstream represents a main obstacle for miRNA-based drugs in vivo and it is considered the fourth main challenge [76]. To evade the degradation by RNase or components of the immune system present in body fluids, the binding and embedding of miRNA molecules with carrier molecules or particles are essential.

On the other hand, CRISPR/Cas system revolutionized miRNA-based cancer therapy by making it possible to quickly alter oncogenic miRNA genes in living cells and animal studies. For instance, CRISPR-based techniques lead to a decrease of several miRNA expressions with high effective rate than traditional methods [77]. Similarly, CRISPR/Cas9 successfully knocked down the expression of oncogenic miRNA in different types of cancer including ovarian cancer cells [78] and brain tumor cells [79]. These findings suggest that CRISPR/Cas can decrease cancer cell survival and multiplication based on miRNA gene knockdown.

Furthermore, CRISPR/Cas gene-targeted cancer therapy is entering preclinical trials [80]. The difficult issue of discovering miRNA targets has been approached in a variety of ways, and it can be modified to target a variety of genes in vivo, underlining its significant promise for future therapeutic use.

Innovative advances in CRISPR/Cas miRNA-editing technology

As previously reported [81], targeting miRNA for either stimulation or suppression of gene expression had a significant impact on studying cancer biology and as a tool for cancer prognosis and treatment. However, the techniques currently in use still show several limitations and require further studies to improve their safety and effectiveness.

As a result, it has been shown that CRISPR/Cas technology provides a promising new therapeutic approach for miRNA targeting, especially when used as an inhibitor. For example, knocking out miRNA-155 by CRISPR/Cas9 in macrophage cell lines shows a great reduction in the development of rheumatoid arthritis-related symptoms [82]. Moreover, CRISPR/Cas9 shows less stimulation of proinflammatory cytokines when compared to siRNA knockout methods, resulting in higher safety for in vivo application [82, 83].

Several studies have investigated CRISPR/Cas-mediated knockdown of miRNA genes and miRNA transcripts in vivo and in vitro (Table 2 [77, 82, 84,85,86,87,88,89,90,91,92,93,94,95]). For example, Chang et al. [77] performed CRISPR/Cas9 knockdown on three miRNA genes expressed in two different colorectal cancer cell lines, HCT116 and HT-29. They showed that CRISPR-based techniques lead to a decrease in miRNA-141, miRNA-17, and miRNA-200c expression with an effectiveness rate of 96% higher than traditional control vectors. Similarly, Huo et al. [78] used lentiviral CRISPR/Cas9 constructs to successfully knock down the expression of pre-miR-21 in ovarian cancer cells. This led to the up-regulation of the miR-21 target genes PDCD4 and SPRY2, which inhibited the growth, migration, and invasion of ovarian cancer cells. In addition, El Fatimy et al. [79] demonstrated that down-regulation of miR-10b led to reduced miR-10b levels in glioma and brain tumor cells. These findings provide evidence for the potential of the CRISPR/Cas technology to inhibit the survival and multiplication of cancer cells, as well as down-regulate their expression.

Table 2 Targeting miRNA genes and miRNA transcripts by CRISPR/Cas in vivo and in vitro studies

CRISPR/Cas-based miRNA gene editing

DNA can be added, removed, or changed at endogenous loci in a genome using the sophisticated genetic engineering technique known as genome editing with sequence-specific nucleases [96]. To produce double-strand breaks (DSBs) on target DNA at a predetermined locus, sequence specific nucleases (SSNs) act as molecular scissors. Endogenous DNA repair processes are triggered in cells by the creation of DSBs. Non-homologous end joining (NHEJ), the predominant DNA repair process in higher eukaryotes, is very error-prone and can result in insertions, deletions, or mismatch alterations at targeted loci. In contrast to NHEJ, homology-directed repair (HDR) needs a DNA donor template and is less efficient [97]. Genome editing has recently become popular due to the development of zinc finger nuclease (ZFN) [98], transcription activator-like effector nucleases (TALEN) [99], CRISPR-Cpf [100], and CRISPR-Cas9 technologies [101].

The first technology to provide a Cas effector for precise genome editing was the DNA-targeting Cas9 system from CRISPR-class 2. Cas9 [102] and Cas12 [103] are two examples of class 2 putative enzymes that have been found as programmable miRNA gene-targeting modules for biotechnological applications due to their widespread use in genome engineering across a variety of species. Both Cas9 and Cas12 use analogous mechanisms to detect and break the complementary DNA sequence that corresponds to the RNA guide [104].

Nuclease activity in these RNP effectors is triggered by the identification of a specific sequence of DNA. The RNP or Cas protein scans long DNA sequences and recognizes the PAM region, then, in turn, initiates an ATP-independent DNA unwinding process and finally results in the pairing of the DNA target strand (TS) and the RNA guide [105]. During the RNA–DNA hybridization process, the “non-target” DNA strand becomes unpaired from the targeted strand, and the Mg2+-dependent endonuclease exploits two active sites (HNH and RuvC in Cas9) or a single active site (RuvC in Cas12) to cleave DNA strand independently (Fig. 4) [101].

Fig. 4
figure 4

The CRISPR/Cas system and gene editing mechanism. a CRISPR/Cas locus in the bacterial genome with associated transcription and translation products. b CRISPR/Cas engineered for site-specific gene editing. c Editing of dsDNA using CRISPR/Cas12a and CRISPR/Cas9 respectively. CRISPR/Cas clustered regularly interspaced short palindromic repeats/CRISPR-associated protein, dsDNA double strands DNA, crRNA CRISPR RNA, TracrRNA trans-activating CRISPR RNA, sgRNA single-guide RNA, PAM protospacer adjacent motif

Cas9-based mechanism for miRNA gene targeting

Large multidomain CRISPR/Cas9 enzymes of type II vary in size from 700 amino acid residues (subtype II-D) to over 1700 amino acid residues (subtype II-C) [106]. Cas9 is part of a functioning RNP that also includes either a crRNA-tracrRNA scaffold hybrid or an engineered crRNA-tracrRNA fusion sgRNA [107]. Cas9’s bilobed structure, made up of a CRISPR RNA-targeted strand (crRNA-TS) pair recognition lobe (Rec lobe) and a nuclease lobe (Nuc lobe), is responsible for target binding and cutting [108, 109]. Rec1, Rec2, and Rec3 are all subdomains of the Rec lobe, while the Nuc lobe has RuvC, HNH, and wedge-PAM-interacting areas as its subdomains.

For PAM identification, apo-Cas9 must first change from its open state when no gRNA is present to its closed state after guide recruitment [110, 111]. Together with the Rec lobe, the RuvC and HNH nuclease domains change their respective conformations upon target DNA binding to activate the nuclease, resulting in a blunt-ended DNA cut [112].

To find potential sequence targets, Cas9 either uses guided 1D diffusion or random 3D collisions to detect PAM sequences along DNA sequences [113]. DNA treated with N4-cystamine forms disulfide bonds with Streptococcus pyogenes that have been modified with cysteine. Modification of Cas9 (T1337C) allowed the temporary interrogation state to be captured and contributed to the understanding of how Cas9 “reads” DNA [114]. Interestingly, by binding to PAM, Cas9 aggressively bends and twists DNA, which causes the nucleotides to flip out of the duplex and toward the gRNA, allowing for interrogation miRNA gene sequence.

The gRNA contains a short region of RNA that is complementary to the target DNA sequence, which allows the gRNA to hybridize with the DNA and form a stable duplex [115]. This RNA–DNA hybridization is specific, meaning that the gRNA will only bind to a target DNA sequence that has a complementary sequence to the gRNA (Fig. 5).

Fig. 5
figure 5

An illustration shows how Cas9 stabilizes the R-loop. In order to construct an R-loop structure, the targeted strand of DNA must link with the 20-nt of gRNA as the DNA curves. The sgRNA not matched to a target DNA sequence (0-nt RNA: DNA) that has a complementary sequence to the gRNA. DNA binding at a sequence matching the 20-nt sgRNA helps proteins accept the RNA–DNA helix and displaced non-target DNA strand. Complete R-loop formation constitutes the signal for the subsequent structure of the targeted gene. gRNA guide RNA, nt nucleotide, sgRNA single-guide RNA, Cas9 CRISPR-associated protein 9, PAM protospacer adjacent motif, Nuc nuclease, Rec recognition

As the DNA bends, the targeted strand pairs with the 20-nt of gRNA to form an R-loop structure [116]. DNA binding at a sequence that matches the 20-nt sgRNA enhances structural changes in proteins that make accommodation for RNA–DNA helix and the displaced non-target DNA strand. Cas9 stabilizes the R-loop by interacting with both ends of the open DNA helix, creating a structural distortion of 30° in the helical bend angle [115]. After the initial melting of the DNA, the gRNA-TS gradually pairs to make a 10-base pair heteroduplex. Crystallization and X-ray analysis of a 10-nt match complex at a resolution of 2.8 Å demonstrate that the TS and non-target strand (NTS) maintain their hybrid state at the PAM-distal end of the DNA platform [112].

Variation in miRNA genes will alter the biogenesis of the miRNA and makes a significant impact on the miRNA transcription, maturation, and target specificity [117]. Thus, gene-editing approaches like the CRISPR/Cas system, which can alter genomic sequences, offer an opportunity for the management and therapy of a wide range of diseases and disorders. For example, Zhou et al. [118] showed that using CRISPR/Cas9 to target mutations in miRNA genes of rice is very effective and is one of the most dependable approaches to finding mutations. Additionally, to silence miRNA genes, Zhao et al. [119] demonstrated that Cas9 could be guided to the target strand by creating DSBs using different gRNAs. They successfully silenced miRNA genes, including those encoding carcinogenic miRNAs like miR-21 and miR-30a, by applying the CRISPR/Cas9 system.

Direct transcripts of the miRNA gene, or pri-miRNA, are subsequently converted into short mature miRNA by Drosha and Dicer enzymes [120]. CRISPR/Cas9 can cleave miRNA Drosha and Dicer processing sites, at the DNA level, which results in indels of varying sizes in the miRNA gene sequences in vivo and in vitro [121, 122]. Cells transfected with the indicated CRISPR/Cas9 constructs reduced mature miR-17, miR-200c, and miR-141 expression by almost 96% relative to control vector-transfected cells [77]. Additionally, clones with only two deletions (GT) or one insertion of nucleotides (A-T) can obstruct the exogenous expression of the miR-17 gene in addition to single clones with substantial portions of deletion (such as 6–18 bp) that prevent mature miR-17 gene synthesis.

These findings provide significant evidence for the theory that CRISPR/Cas9-induced alterations in the stem-loop structure of pri-miRNA can inhibit the synthesis of mature miRNA.

Cas12-based mechanism for miRNA gene targeting

Cas12 proteins, in contrast to Cas9, are large multidomain enzymes with a wide range of diversity that has led to their classification into more than ten distinct subclasses such as V-A to V-K, and other subclasses V-U [106, 123]. They vary in the processes via which they form RNPs, the RNPs themselves, and the nucleic acids to which they bind. RNPs can be functional when guided by either a single crRNA [124] or a hybrid of a crRNA and a trans-activating CRISPR RNA (tracrRNA) [125] or a crRNA and short-complementarity untranslated RNA (scoutRNA) [126]. Generally, Rec and Nuc lobes, similar to those found in Cas9, are present in even the smallest Cas12 effectors, like Cas14 (referred as Cas12f; subtype V-F) [127, 128], Cas12j (subtype V-J) [129] and Cas12g (subtype V-G) [130]. Beyond the similar bilobed structure, structural analyzes of different Cas12 proteins revealed remarkable homogeneity. Specifically, a crRNA oligonucleotide-binding domain and the RuvC domain give rise to the Rec lobe domains (Rec1 and Rec2) and the DNA-loading “nuclease” (Nuc) or zinc-ribbon domains, which together provide a highly adaptable platform. Other short domains are sometimes fused or added into this basic structure to facilitate PAM identification and NTS binding or to direct recruiting efforts via a zinc-finger motif.

Cas12a is an example of a Cas12 protein that uses a method similar to that used by other Cas12 proteins for binding and cutting DNA. Cas12a forms an adaptable “open” conformation in the absence of a gRNA, and a “closed” conformation that is prepared for PAM recognition upon crRNA binding [131, 132]. The RuvC active site for nuclease suppression is structurally occluded by the Rec domains in the closed conformation [131]. Rec domains rearrange to make space for the heteroduplex form when a dsDNA target is unwound in a PAM-dependent manner and then hybridizes with a crRNA [133]. This conformational change occurs simultaneously with the activation and opening of the RuvC active site, which successively cleaves the single-stranded NTS and then the TS to produce a 5′-overhang staggered DNA DSB [134, 135]. Recent structural analysis of Cas12i and Cas12j’s active sites found that the single strand DNA (ssDNA) substrate coupled with two magnesium cofactors provides mechanistic detail for the two-metal ion catalysis process of Cas12 proteins [136]. Nonspecific ssDNA “shredding” in trans remains active in the RuvC domain after cis DNA cutting [137]. The structural details of DNA interrogation are still unclear, although it is speculated that DNA bending is involved in the duplex opening after the PAM [138]. Data showing that DNA distortion can reveal ssDNA segments, which are then identified by Cas12, provide validity to this mechanism [139].

In particular, the PAM-interacting domain has been observed frequently close to unraveled DNA and may play a key role in this process. However, it is unclear whether these domains and other nearby components effectively participate in DNA unwinding or only connect with the unwound DNA for stability. To advance the design of Cas12-based genome editing tools, a comprehensive understanding of the DNA interrogation mechanism is needed.

Regions of miRNAs genes targeted with CRISPR/Cas

Genomic engineering tools like the CRISPR/Cas system have been used to target different parts of the post-transcriptional miRNA sequences at the DNA level for therapeutic and diagnostic purposes. These parts include the Drosha cleavage site, the 5ʹ-end terminal, the upstream region of the miRNA, the secondary stem-loop structure, and/or the mature miRNA (Fig. 6). More critically, according to the Jiang et al. [140] study, which used CRISPR/Cas9 to knockout miR-93 in HeLa cells, deletion of a single nucleotide in a pre-miRNA DNA region can result in the highly specific total knockout of the targeted miRNA.

Fig. 6
figure 6

An illustration showed different targeting DNA regions of the post-transcriptional editing of miRNAs with the CRISPR/Cas system. Targeting terminal loops: CRISPR/Cas9 has the potential to inhibit the production of monoisotopic miRNAs by targeting either the terminal loop or the 5ʹ region of the pre-miRNA. Targeting secondary stem loop: Random targeting of pri-miRNA at secondary stem-loop structure by CRISPR/Cas9. Targeting mature miRNA: CRISPR/Cas9 sequences that target mature miRNA are used to successfully inhibit miRNA expression. Dicer cleavage sites: The expression levels of mature miRNAs can be successfully down-regulated by gRNA when it has been directed to target miRNA in the Dicer site. Drosha cleavage sites: gRNA successfully targets mature miRNAs at the Drosha region to bring down the expression levels of certain miRNAs. CRISPR/Cas clustered regularly interspaced short palindromic repeats/CRISPR-associated protein, gRNA guide RNA, sgRNA single-guide RNA, crRNA CRISPR RNA

Interestingly, the CRISPR/Cas technology can be used to specifically target the terminal loop or 5ʹ region of the pre-miRNAs and inhibit their biogenesis. By using CRISPR/Cas9 to target a specific location in miR-93 in HeLa cell line, Jiang et al. [140] showed that targeting terminal loop or 5ʹ region of the pre-miRNA at the DNA level is one of the efficient ways to impair the RNA biogenesis in monoisotopic miRNA. Furthermore, they confirmed the functional knockdown by quantitative PCR analysis.

Similarly, the CRISPR/Cas technology can be used to specifically target miRNA loops and seed region of the pre-miRNAs and inhibit their biogenesis (Fig. 6). The stem-loop structure of pre-miRNAs exposes a duplex of the mature miRNA on the stem, where it can be cleaved off by the deoxyribonuclease Dicer-like 1 (DCL1) [141]. In plants [142], the pri-miRNA is turned into the miRNA duplex in the nucleus. In animals [143], this step takes place in the cytoplasm. In vitro, experiments by Chang et al. [77] showed that CRISPR/Cas9 random targeting of pri-miRNA to the secondary stem-loop structure effectively reduced miRNA families’ expression. Furthermore, they showed that the production of mature miRNA can be inhibited by introducing mutations into the stem-loop DNA structure of pri-miRNA using the CRISPR/Cas9 system.

Moreover, the CRISPR/Cas technology allows for the exact targeting of miRNA seed DNA sequences, which can then be utilized to inhibit the functions of miRNAs. The first 2–8 DNA nucleotides of a miRNA, counting from the 5ʹ end to the 3ʹ end, are known as the seed sequence, and this area has been conserved in a heptametric pattern [144]. Seed-mediated target recognition is the primary method by which miRNAs suppress translation or cause mRNA de-adenylation or disintegration [145]. Thus, the scientist tried to find out the impact of targeting the seed region of miRNA by CRISPR/Cas system to see the impact of the change on the miRNA function. For example, one of the studies performed by the Jiang et al. [140] revealed that CRISPR/Cas9 can be used as a proper tool for knocking out the miRNA-93 at the DNA level. They used HeLa cells and designed a specific gRNA to target the PAM sequence of miRNA-93 in the 5ʹ region which is one of the critical onco-miRNA in cancer prognosis [140]. They demonstrated that minor indels in the 5ʹ region of this miRNA cause sequence impairment and biogenesis inhibition, resulting in an effective and selective gene knockout.

Additionally, the CRISPR/Cas technology can be used to specifically target mature miRNAs in DNA and suppress their expression. For example, the successful silencing of miRNA expression by targeting mature miRNAs was obtained by Chung et al. [146] while aiming to shut down the expression of genes involved in drought response in rice. The team showed that targeting of mature miRNA sequences, as well as of miRNA biogenesis sites through CRISPR/Cas9 technology is a powerful tool for studying loss of function mutation in miRNAs in rice.

Further, it was demonstrated conclusively that mutations introduced by CRISPR/Cas at miRNA Drosha- and Dicer-processing sites lead to the down-regulation of mature miRNA by blocking biogenesis. CRISPR/Cas9 system can target Drosha and Dicer cleavage sites of the pri-miRNAs and inhibit their biogenesis [77]. The complex process of miRNA maturation requires the action of numerous Drosha and Dicer enzymes, which cleave particular locations in the pri-miRNA sequence to produce the complete and functional miRNA structure. The flanking and internal structures of the pri-miRNA, which determine the effectiveness of these cutting enzymes, have an impact on the specific cutting site [147]. Chang et al. [77] created a gRNA to target mature miR-17, miR-200c, and miR-141 in Dicer and Drosha cleavage sites at the DNA level because of their critical function in the biogenesis of miRNA. By transfecting cells with specific CRISPR/Cas9 constructs, they were able to reduce the production of these mature miRNAs by as much as 96% compared to cells transfected with control vectors. The results of the above study added credibility to the argument that the CRISPR/Cas9 system is highly specific for editing miRNA sequences, as it can avoid off-target effects even when modifying miRNAs within the same family or with highly conserved regions.

Advantages of CRISPR/Cas miRNA targeting

CRISPR/Cas is a more effective, precise, and stable technique for activating or silencing miRNAs in cancer cells than recent methods. For example, the effects of an inserted mutation can be permanently integrated into the genome and passed on to the next generation of cells, which is a major advantage of using genetic engineering, especially the CRISPR/Cas system. According to Friedland et al. [148] study on Caenorhabditis elegans, mutations created by CRISPR/Cas9 can pass to the offspring. Likewise, according to Chang et al. [77], who cloned CRISPR/Cas9 designs with sgRNAs that target the biogenesis processing sites of certain miRNAs, CRISPR/Cas9 can substantially and specifically suppress the production of these miRNAs by up to 96%. They also demonstrated that similar results could also be obtained in vivo by transfecting mice with the CRISPR/Cas9 construct and targeting miRNA-17.

Furthermore, when compared to other methods, CRISPR/Cas offers higher accuracy and specificity (Table 3). For instance, the findings of Wu et al. [149] demonstrated that technologies based on CRISPR had great accuracy and specificity in comparison to other techniques. This precision is a result of the CRISPR/Cas system’s focus on sgRNA, which is targeted to a specific site throughout the entire genome. Moreover, Chung et al. [146] utilized CRISPR with recombinant codon-optimized Cas9 (rCas9) on 13 different types of miRNA in rice, obtaining a 59.4% rate of mutations, including mono-allelic (8.54%), bi-allelic (11.1%), and hetero-allelic combination (39.7%) mutations. Similarly, Narayanan et al. [85] proved that the CRISPR/Cas9 approach for miRNA gene knockdown can reach a high success rate in mammalian cells with minimal off-target activity. Further, they found that in mammalian cells, the CRISPR/Cas9 method for miRNA gene knockdown can achieve a success rate of 75–99% with low off-target activity. In addition, the effectiveness of miRNA silencing through CRISPR/Cas9 genome editing is also enhanced by the ability of this technique to target not only a single miRNA sequence but also several pre-miRNA structures in a single application [150].

Table 3 Comparison between the five most common genome editing tools

In addition to precision and selectivity, the CRISPR/Cas system allows for the targeting of various loci within the miRNA gene. Clinical trials are continuing to explore new medicines to fight diseases in humans using a wide variety of Cas9 nuclease variants [151]. However, Cas9’s effectiveness and adaptability in genome editing are capped by its toxicity and potential for mutagenicity [152, 153]. In response to these challenges, researchers have enhanced their pursuit of CRISPR/Cas systems, which have the potential to be refined into next-generation genome editing tools. Cas12a, Cas12b, Cas12f, Cas12g, and Cas14 are some of the newest type V members in class 2 to receive scientific attention [130, 154,155,156]. For genome editing, Cas12a is interesting because it appears to be more precise than Cas9 [157] and has opportunities for future development [158, 159]. Cas12a’s pre-crRNA is processed into mature crRNAs due to its unique RNase activity by the proper Cas12a [90, 160], and its short crRNA (40 nt) helps overcome the size challenge in delivery via viral vectors [161]. As a result, through the delivery of numerous crRNAs on a single plasmid, this Cas12a activity has been effectively used in multiplex gene regulation to successfully edit many endogenous targets at once [162, 163].

In light of this, the CRISPR/Cas system has been widely used as a genetic engineering tool in a variety of animals, as well as in vivo and in vitro studies of human disorders, by taking advantage of the endogenous DNA repair machinery of cells.

Challenges of targeting miRNAs by CRISPR/Cas and its strategies to overcome

Cas9 PAM deficiency in miRNA sequence

CRISPR gRNA is composed of two main parts, crRNA and tracrRNA. The crRNA sequence targets and binds to the NGG, PAM, and the target DNA, whereas tracrRNA has a role in ensuring Cas9 nuclease activity [164]. Requiring a PAM sequence next to the target site is considered one of the main limitations of using the CRISPR/Cas9 system for targeting miRNA genes at the DNA level [165]. As miRNAs are very short sequences of nucleotides, often it is quite challenging for these small fragments to find the PAM region of the targeted miRNA. Moreover, according to the Bi et al. [166] study, most of the miRNA sequences do not contain the classic mammalian 5ʹ-NGG-3ʹ PAM region, which is essential for Cas9 protein activity. And this issue becomes even more obvious when anyone tries to target mature miRNA which is very short, usually around 20–24 nucleotides long.

To overcome this issue, advanced bioinformatics platforms can be used to assess the presence or lack of an appropriate PAM region in the target miRNA sequence. Zhou et al. [118] found that out of the 592 rice miRNAs assessed through the miRBase portal (http://www.mirbase.org/), 556 miRNAs (93.92%) presented a suitable PAM site for Cas9, showing the importance of these bioinformatics tools in successful planning experiments.

Further, additional approaches to get around this limitation include using different Cas proteins, including Cas12a, which potentially overcome this challenge. Cas12a is a single RNA-guided endonuclease, which means it processes its own gRNAs and only needs crRNA for targeting [167]. For instance, miR-21 can reprogram microglial cells and establish a favorable environment for cancer development in glioma cells [168]. It affects native brain cell types such as endothelial cells, neurons, as well as invading monocytes and macrophages which make up-regulation of cytokine productions [169]. As a result of its ability to recognize the PAM sequences in miR-21, CRISPR/Cas12a can reduce miR-21 expression by altering the coding sequences and regulating cell proliferation in vivo and in vitro [170] (Fig. 7).

Fig. 7
figure 7

An illustration shows applying Ca12a instead of Cas9 protein, which has the ability to recognize PAM sequences in miRNA editing. By altering the miR-21 coding sequences in glioma cells, CRISPR/Cas12a decreases miR-21 expression through microenvironment cells, which controls both in vitro and in vivo cell proliferation. CRISPR clustered regularly interspaced short palindromic repeats, PAM protospacer adjacent motif, sgRNA single-guide RNA

Besides that, the PAM sequence required for Cas12 is "TTTV" (where V is A, C, or G), but the PAM sequence needed for Cas9 is "NGG" (where N is A, C, T, or G) [105]. Nevertheless, scientists have found a unique property of Cas12 that can be used for something other than editing the genome such as cutting ssDNA. These recently uncovered features of Cas12 make CRISPR an attractive new area for targeting miRNAs in cancer therapy.

Furthermore, multiple Cas proteins, as shown in Table 4 [107, 141, 154, 171,172,173,174,175,176,177,178,179], bind to various PAM sites. Therefore, the success rate of CRISPR-based miRNA editing could be improved by using a variety of Cas proteins as well.

Table 4 List of Cas proteins with their PAM sequence targets

Short sequence of miRNA

One drawback of CRISPR gene editing is the challenge of detecting sgRNAs besides PAM, particularly when looking at miRNAs, which have a short nucleotide sequence of 20 to 24 bases. The same issue is also present when using other inhibitors, such as antisense RNAs, to target miRNA. Moreover, not all sgRNA is efficient and working properly. Thus, sometimes a PAM sequence is available in a miRNA gene, the sgRNA might not be efficient enough to knock down the miRNA.

To design a reliable and effective sgRNA, it is important to use properly in silico techniques to show all the predictions [180]. Another way to overcome this issue is to use more than one Cas protein to target different parts of the interested gene. For instance, Godden et al. [87] showed that when two sgRNAs are used together, the targeted miRNA gene loses all of its functions in embryos. Additionally, the discovery of new and more Cas nucleases with broader PAM recognition sequences gives more opportunities to design gRNAs in different parts of miRNA sequences [181].

Off-target effects in CRISPR/Cas-mediated miRNA editing

Off-targeting, which results from a gRNA mismatch binding to the wrong target, is one of the key challenges of using the CRISPR/Cas system [182]. All gene silencing approaches, including the CRISPR/Cas system, is characterized by off-target effects, which have high frequency [183]. Basically, off-targeting happens when the gRNA binds to the wrong target due to similarities between different sequences of the same genome, especially in mammals, and it might lead to further mutation and disorders [5].

CRISPR has attracted a lot of attention as a potential new gene editing tool over the past two decades. However, it has proven of limited use due to its tendency to be off-target. Because of this, a growing number of studies have spent the last few years working to enhance the system’s editing abilities while also reducing the number of undesired consequences. Include in particular the following strategies to overcome off-targeting:

Firstly, increasing sgRNA specificity. The specificity of sgRNA and target DNA recognition is determined by the number of base pairs in the region of the sgRNA, typically 10–12 bp [149]. The off-target effects are also modified by the remaining sequence to various levels. The efficacy of sgRNA-based gene editing was found to increase in direct correlation with the GC level of the seed region. The off-target effect is reduced or disappears when there are three or more base mismatches between the sgRNA seed region sequence and the DNA sequence at the off-target position [184]. Accordingly, 40–60% of the GC content can be used during sgRNA sequence design. The specificity of sgRNA can be increased by using sequences that are highly dissimilar to those of the off-target genes [184].

In addition, the specificity of a sgRNA is highly linked to its length. Using shorter sgRNA sequences (those with less than 20-nt) has been shown to decrease the off-target effects without reducing the efficiency of gene editing [185]. Shortening sgRNA sequences, however, may not improve specificity, and may even decrease gene editing efficiency. Therefore, more studies are needed to confirm the effectiveness of the method of decreasing the off-target effect by decreasing sgRNA length. Furthermore, adding two guanines (called ggX20 sgRNAs) to the 5ʹ end of a sgRNA in place of the matching GX19 sgRNAs during the design process is another simple alteration that may be made to boost the specificity of the sgRNA and decrease its off-target effect [183, 186]. Off-targeting can also be reduced by using sequence gRNA design and off-target evaluation sites that are available online (Table 5 [187,188,189,190,191,192,193,194,195,196,197,198,199,200,201]).

Table 5 List of main Bioinformatics tools with their features and types of Cas proteins

Secondly, Cas-sgRNA dosage control. One of the other strategies that can be used to minimize the consequences of off-target effects is controlling the amount of sgRNA or Cas protein produced. For example, when Pattanayak et al. [202] cut missed loci in the HEK293T cells’ genomes, they discovered that short sequences, low sgRNA activity compared to sequence length, high activity has better specificity, and high amounts of sgRNA-Cas9 compounds can cut close internal sites or PAM sequences. Cas9, which has DSB activity, is continuously expressed in cells, which could enhance the risk of off-target sequences. This risk can be minimized by blocking antibodies or inhibitors of the Cas9 protein, which limits the off-target effect. Regulating the concentration of sgRNA and Cas9 nucleases can lower the off-target risk, but this will also reduce the corresponding genome editing ability. Thus, it is necessary to balance the efficiency of gene editing against the possibility of producing an off-target effect. Recent studies proved that decreasing the Cas9:gRNA complex ratio to 1:2 or 1:3 improves knockout efficacy and successfully decreases off-target effects [108, 203].

The third strategy is the chemical modification of sgRNA. Besides the current strategies, chemical modification of sgRNA is also an effective way to minimize off-target effects. Chemical modifications to crRNA include 2-fluoro ribose, and 2-O-methyl-3ʹ-thiophosphate (MS), which can strengthen the selectivity of Cas9 endonuclease and the stability of sgRNA [204, 205]. A further factor crucial to successful gene manipulation is the strategy for optimizing off-target effect identification. Off-target effects of CRISPR/Cas9 can be detected using tools like integration-deficient lentiviral vectors (IDLVs) containing integrase defects, and this method can identify an off-target frequency of at least 1% [206].

Targeting multi-sites of gene by one miRNA or targeting one site/pathway by multiple miRNAs

As discussed in the previous section, miRNAs undergo a complex molecular pathway within the cell, which is not fully elucidated. According to “seed sequence matching” bioinformatics research, a single miRNA can control hundreds of target genes, while many miRNAs can trigger a single gene. For example, the miR-17-92 cluster suppresses cyclin-dependent kinase inhibitor 1A (CDKN1A), E2F transcription factor 1 (E2F1), and PTEN, which can cause up-regulated cell growth; miR-200 targets ZEBs to induce E-cadherin, which inhibits EMT [77]. Therefore, there is a potential for different consequences when researchers attempt to target a single miRNA to reduce the amount of certain proteins by focusing on the miRNA that complements its mRNA.

Conversely, the protein-synthesis pathway of a gene can be regulated by more than one miRNA [207]. For example, using a high-throughput luciferase reporter screen, Wu et al. [208] found that 28 miRNAs can directly reduce p21Cip1/Waf1 or CDKN1A by targeting its 3ʹ untranslated region. Furthermore, many of these miRNAs were found to be elevated in malignancies, suggesting that they could act as oncogenesis modulators.

To obtain a complete therapeutic effect, targeting just one miRNA region in the genome is not enough. As shown in Fig. 8, one miRNA or more than one miRNA can target the same gene, which means that targeting just one miRNA is not enough to interrupt a disease as long as more than one miRNA controls the same gene. Therefore, to overcome this challenge, it is best to create or develop multiple gRNA that can be utilized to target different genomic loci or mature miRNA genomic sequences (Fig. 8). Kabadi et al. [209] previously developed a single lentiviral strategy to express and deliver a Cas9 nuclease and three to four sgRNAs transcribed from separate and distinct RNA polymerase III promoters. Interestingly, multiple gene editing and prolonged transcriptional activation were facilitated by the high levels of expression of individual sgRNAs in HEK293T and primary human dermal fibroblasts cells. This indicates the potential utility of this approach in miRNA-based biomedicine [209]. Moreover, there are different vectors with different capacities, such as adenovirus vectors, which can successfully express eight different multiplex gRNAs [210]. This strategy holds promise for minimizing the number of CRISPR/Cas components used and the risk of undesirable side effects of the co-delivery of several sgRNAs encoding constructs.

Fig. 8
figure 8

An illustration shows the main strategy to overcome multiple sites targeting or one site targeting through the CRISPR/Cas system. The CRISPR/Cas system targets one miRNA, which in turn suppresses miRNA to bind numerous genes and limit the synthesis of tumor protein. The CRISPR/Cas system is designed to target numerous miRNAs, which in turn restrict the same oncogenic gene, thereby limiting the growth of tumors. The CRISPR/Cas system is intended to target a large number of miRNAs, which then block the activity of the same oncogenic gene. As a result, the progression of tumor cell is suppressed or slowed down. CRISPR clustered regularly interspaced short palindromic repeats, sgRNA single-guide RNA, Cas CRISPR-associated protein

Delivery challenges of CRISPR/Cas system both in vivo and in vitro

Delivery of CRISPR/Cas presents numerous challenges, such as the selection of a good, safe, and precise vector. Inappropriate vectors are associated with a higher risk of toxicity and off-targeting [5]. Likewise, CRISPR/Cas delivery is far more challenging to get precisely in vivo than in vitro [211]. Toxicity, size capacity, and mismatching are the three main types of in vivo delivery difficulties presented by the CRISPR/Cas system [5, 9, 212, 213].

CRISPR/Cas vectors can be broadly classified into two types: viral and non-viral [214]. Interestingly, an increasing number of studies are selecting lentivirus as their vector of choice. However, using a viral vector in vivo has a number of drawbacks, including insertional restriction, immune response, and size capacity (Table 6 [78, 87, 88, 119, 215,216,217,218,219,220,221,222]). Moreover, the risks of off-targeting and further mutation rise with prolonged expansion following insertion [223].

Table 6 Summarizes the most common delivery approaches that can be used for both in vivo and in vitro CRISPR/Cas system delivery

Strategy to overcome delivery challenges

Vector capacity is a significant limitation during CRISPR/Cas delivery, especially when viral vectors such as adeno-associated virus (AAV) are used. However, several strategies can be used to overcome or minimize this problem, like picking up a smaller Cas protein or using two vectors instead of one. Several Cas proteins have been categorized or selected according to their molecular weight; for example, a smaller version of Cas9 was discovered in Staphylococcus aureus, showing that size optimizations have been performed. This Cas9 variant is 1 kb smaller than the original Cas9 from Streptococcus pyogenes; therefore, it may be integrated into a single AAV vector [224]. Furthermore, Cas14 has a two-fold smaller molecular size than Cas9 [225]. Secondly, more than one viral vector can be used to deliver the CRISPR/Cas system [9, 226]. For example, using two vectors rather than one reduces the off-targeting risk, which rises in parallel with the vector’s size [5, 224].

On the other hand, non-viral vectors are delivery systems that can be used to transport therapeutic molecules, such as CRISPR/Cas systems, into target cells without the use of viral vectors in both in vivo and in vitro [223]. The advantages of using non-viral vectors for CRISPR/Cas delivery into cancer cells include their safety, low immunogenicity, ease of preparation, versatility, minimal off-targeting, and less exposure to nuclease [223, 227, 228]. However, non-viral vectors have some crucial drawbacks such as degradation in vivo experiments, varied biocompatibility and toxicity, low delivery efficiency, and restricted delivery efficiency [228].

Several types of non-viral vectors can be used for CRISPR/Cas delivery, including liposomes, polymeric nanoparticles, and viral-like particles. Liposomes are spherical vesicles composed of a lipid bilayer that can be used to encapsulate nucleic acids, including CRISPR/Cas components [229]. Polymeric nanoparticles are composed of synthetic polymers that can also be used to encapsulate CRISPR/Cas components [230]. Finally, viral-like particles are self-assembling protein cages that can be used to deliver CRISPR/Cas components [231].

Conclusions and perspectives

The current study underlines the background of the CRISPR/Cas system in miRNA-based cancer therapy. The CRISPR/Cas system provides new insights into cancer therapeutics that were previously unexplored in our understanding of the non-coding genome. With the development of CRISPR/Cas-based gene editing technology, it is now possible to target mutations in a precise and permanent way. Short non-coding RNAs like miRNA can also be targeted in a precise way at the DNA level. Therapeutic genome editing based on CRISPR/Cas-miRNA targeting is moving from preliminary research to preclinical development. The challenging task of identifying miRNA targets has been approached in a number of ways, including the application of CRISPR screening and miRNA gene alteration. When there is a miRNA mutant, CRISPR knockout libraries can be used to find target genes whose mutation fixes the miRNA mutant phenotype. Interestingly, the biological role of specific sites can be explored through specific sgRNA libraries that target miRNA binding sites. Custom libraries can be delivered into wild-type cells to select cells with binding site mutations that mimic the oncogenic miRNA. The CRISPR/Cas technology is currently undergoing clinical trials for the treatment of cancer, and its application in cancer immunotherapy and the inactivation of cancer-causing viral infections holds promise for addressing altered cancer cells and extending the scope of cancer therapeutic targets based on miRNA therapy. Currently, the potential use of CRISPR/Cas as a miRNA targeting platform in cancer therapy has only been partially explored, and it needs further studies.

Availability of data and materials

The datasets generated and analyzed during the current study are available from the corresponding author on reasonable request.

Abbreviations

AAV:

Adeno-associated virus

ADAR:

Adenosine deaminase RNA specific

AGO:

Argonaute

CDKN1A:

Cyclin-dependent kinase inhibitor 1A

CRISPR:

Clustered regulatory interspaced short palindromic repeats

crRNA-TS:

CRISPR RNA-targeted sequence

CYP:

Cytochrome P450

CYP7A1:

Cholesterol 7 alpha-hydroxylase

DCL1:

Dicer-like 1

DSBs:

Double strand breaks

E2F1:

E2F transcription factor 1

EGFR:

Epidermal growth factor receptor

EMT:

Epithelial-to-mesenchymal transition

gRNA:

Guide RNA

HDR:

Homology-directed repair

IDLVs:

Integration-deficient lentiviral vectors

NF-κB:

Nuclear factor-κB

NHEJ:

Non-homologous end joining

NTS:

Non-target strand

Nuc lobe:

Nuclease lobe

ORF:

Open reading frame

PAM:

Protospacer adjacent motif

PIK3R2:

Phosphatidylinositol 3-kinase regulatory subunit 2

Pol II:

Polymerase II

pre-crRNAs:

Precursor CRISPR-RNA

pri-miRNA:

Primary miRNA

PTEN:

Phosphatase and tensin homolog

Rec lobe:

Recognition lobe

RISC:

RNA-induced silencing complex

RNP:

Ribonucleoprotein

scoutRNA:

Short-complementarity untranslated RNA

sgRNA:

Single-guide RNA

SOCS1:

Suppressor of cytokine signaling 1

ssDNA:

Single strand DNA

SSNs:

Sequence specific nucleases

TALEN:

Transcription activator-like effector nucleases

TracrRNA:

Trans-activating CRISPR RNA

TRBP:

Tar RNA-binding protein

TS:

Target strand

UTR:

Untranslated region

VEGF-A:

Vascular endothelial growth factor A

XPO5:

Exportin 5

ZEB1:

Zinc finger E-box-binding homeobox 1

ZEB2:

Zinc Finger E-box binding homeobox 2

ZEBs:

Zinc finger E-box binding homeoboxs

ZFN:

Zinc finger nuclease

References

  1. Ishino Y, Shinagawa H, Makino K, Amemura M, Nakata A. Nucleotide sequence of the iap gene, responsible for alkaline phosphatase isozyme conversion in Escherichia coli, and identification of the gene product. J Bacteriol. 1987;169(12):5429–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. 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(1):273–97.

    Article  CAS  PubMed  Google Scholar 

  3. Brouns SJ, Jore MM, Lundgren M, Westra ER, Slijkhuis RJ, Snijders AP, et al. Small CRISPR RNAs guide antiviral defense in prokaryotes. Science. 2008;321(5891):960–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Makarova KS, Wolf YI, Alkhnbashi OS, Costa F, Shah SA, Saunders SJ, et al. An updated evolutionary classification of CRISPR-Cas systems. Nat Rev Microbiol. 2015;13(11):722–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Rasul MF, Hussen BM, Salihi A, Ismael BS, Jalal PJ, Zanichelli A, et al. Strategies to overcome the main challenges of the use of CRISPR/Cas9 as a replacement for cancer therapy. Mol Cancer. 2022;21(1):64.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science. 2012;337(6096):816–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Mojica FJM, Díez-Villaseñor C, García-Martínez J, Almendros C. Short motif sequences determine the targets of the prokaryotic CRISPR defence system. Microbiology. 2009;155(Pt 3):733–40.

    Article  CAS  PubMed  Google Scholar 

  8. Iacomino G. miRNAs: the road from bench to bedside. Genes. 2023;14(2):314.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Martinez-Lage M, Puig-Serra P, Menendez P, Torres-Ruiz R, Rodriguez-Perales S. CRISPR/Cas9 for cancer therapy: hopes and challenges. Biomedicines. 2018;6(4):105.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Miyoshi J, Zhu Z, Luo A, Toden S, Zhou X, Izumi D, et al. A microRNA-based liquid biopsy signature for the early detection of esophageal squamous cell carcinoma: a retrospective, prospective and multicenter study. Mol Cancer. 2022;21(1):44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Krol J, Loedige I, Filipowicz W. The widespread regulation of microRNA biogenesis, function and decay. Nat Rev Genet. 2010;11(9):597–610.

    Article  CAS  PubMed  Google Scholar 

  12. Peng Y, Croce CM. The role of MicroRNAs in human cancer. Signal Transduct Target Ther. 2016;1:15004.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Hussen BM, Hidayat HJ, Salihi A, Sabir DK, Taheri M, Ghafouri-Fard S. MicroRNA: a signature for cancer progression. Biomed Pharmacother. 2021;138:111528.

    Article  CAS  PubMed  Google Scholar 

  14. Liang L, Cen H, Huang J, Qin A, Xu W, Wang S, et al. The reversion of DNA methylation-induced miRNA silence via biomimetic nanoparticles-mediated gene delivery for efficient lung adenocarcinoma therapy. Mol Cancer. 2022;21(1):186.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Shen P, Yang T, Chen Q, Yuan H, Wu P, Cai B, et al. CircNEIL3 regulatory loop promotes pancreatic ductal adenocarcinoma progression via miRNA sponging and A-to-I RNA-editing. Mol Cancer. 2021;20(1):51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Nombela P, Miguel-López B, Blanco S. The role of m6A, m5C and Ψ RNA modifications in cancer: novel therapeutic opportunities. Mol Cancer. 2021;20(1):18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Taghavipour M, Sadoughi F, Mirzaei H, Yousefi B, Moazzami B, Chaichian S, et al. Apoptotic functions of microRNAs in pathogenesis, diagnosis, and treatment of endometriosis. Cell Biosci. 2020;10(1):12.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Kang M, Tang B, Li J, Zhou Z, Liu K, Wang R, et al. Identification of miPEP133 as a novel tumor-suppressor microprotein encoded by miR-34a pri-miRNA. Mol Cancer. 2020;19(1):143.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Leichter AL, Sullivan MJ, Eccles MR, Chatterjee A. MicroRNA expression patterns and signalling pathways in the development and progression of childhood solid tumours. Mol Cancer. 2017;16(1):15.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Chan SH, Wang LH. Regulation of cancer metastasis by microRNAs. J Biomed Sci. 2015;22(1):9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Bhat AA, Younes SN, Raza SS, Zarif L, Nisar S, Ahmed I, et al. Role of non-coding RNA networks in leukemia progression, metastasis and drug resistance. Mol Cancer. 2020;19(1):57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Wang Y, Wang L, Chen C, Chu X. New insights into the regulatory role of microRNA in tumor angiogenesis and clinical implications. Mol Cancer. 2018;17(1):22.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Dexheimer PJ, Cochella L. MicroRNAs: from mechanism to organism. Front Cell Dev Biol. 2020;8:409.

    Article  PubMed  PubMed Central  Google Scholar 

  24. De Rie D, Abugessaisa I, Alam T, Arner E, Arner P, Ashoor H, et al. An integrated expression atlas of miRNAs and their promoters in human and mouse. Nat Biotechnol. 2017;35(9):872–8.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Kim YK, Kim VN. Processing of intronic microRNAs. Embo J. 2007;26(3):775–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Dastmalchi N, Safaralizadeh R, Banan Khojasteh SM, Sam MR, Latifi-Navid S, Hussen BM, et al. An updated review of the cross-talk between microRNAs and epigenetic factors in cancers. Curr Med Chem. 2021;28(42):8722–32.

    Article  CAS  PubMed  Google Scholar 

  27. Tanzer A, Stadler PF. Molecular evolution of a microRNA cluster. J Mol Biol. 2004;339(2):327–35.

    Article  CAS  PubMed  Google Scholar 

  28. Bohnsack MT, Czaplinski K, Gorlich D. Exportin 5 is a RanGTP-dependent dsRNA-binding protein that mediates nuclear export of pre-miRNAs. RNA. 2004;10(2):185–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Carthew RW, Sontheimer EJ. Origins and mechanisms of miRNAs and siRNAs. Cell. 2009;136(4):642–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Eulalio A, Huntzinger E, Nishihara T, Rehwinkel J, Fauser M, Izaurralde E. Deadenylation is a widespread effect of miRNA regulation. RNA. 2009;15(1):21–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Wilczynska A, Bushell M. The complexity of miRNA-mediated repression. Cell Death Differ. 2015;22(1):22–33.

    Article  CAS  PubMed  Google Scholar 

  32. Xu W, San Lucas A, Wang Z, Liu Y. Identifying microRNA targets in different gene regions. BMC Bioinf. 2014;15:S4.

    Article  Google Scholar 

  33. Valinezhad Orang A, Safaralizadeh R, Kazemzadeh-Bavili M. Mechanisms of miRNA-mediated gene regulation from common downregulation to mRNA-specific upregulation. Int J Genomics. 2014;2014:970607.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Zhang Y, Fan M, Zhang X, Huang F, Wu K, Zhang J, et al. Cellular microRNAs up-regulate transcription via interaction with promoter TATA-box motifs. RNA. 2014;20(12):1878–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Monroig-Bosque PDC, Shah MY, Fu X, Fuentes-Mattei E, Ling H, Ivan C, et al. OncomiR-10b hijacks the small molecule inhibitor linifanib in human cancers. Sci Rep. 2018;8(1):13106.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Fu Z, Wang L, Li S, Chen F, Au-Yeung KK-W, Shi C. MicroRNA as an important target for anticancer drug development. Front Pharmacol. 2021;12:736323.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Vasudevan S, Tong Y, Steitz JA. Switching from repression to activation: microRNAs can up-regulate translation. Science. 2007;318(5858):1931–4.

    Article  CAS  PubMed  Google Scholar 

  38. Friedman RC, Farh KK, Burge CB, Bartel DP. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2009;19(1):92–105.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Iacomino G, Siani A. Role of microRNAs in obesity and obesity-related diseases. Genes Nutr. 2017;12:23.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Lui PY, Jin DY, Stevenson NJ. MicroRNA: master controllers of intracellular signaling pathways. Cell Mol Life Sci. 2015;72(18):3531–42.

    Article  CAS  PubMed  Google Scholar 

  41. Gambari R, Brognara E, Spandidos DA, Fabbri E. Targeting oncomiRNAs and mimicking tumor suppressor miRNAs: new trends in the development of miRNA therapeutic strategies in oncology (Review). Int J Oncol. 2016;49(1):5–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Rhim J, Baek W, Seo Y, Kim JH. From molecular mechanisms to therapeutics: understanding MicroRNA-21 in cancer. Cells. 2022;11:18.

    Article  Google Scholar 

  43. Misso G, Di Martino MT, De Rosa G, Farooqi AA, Lombardi A, Campani V, et al. Mir-34: a new weapon against cancer? Mol Ther Nucleic Acids. 2014;3(9):e195.

    Article  PubMed Central  Google Scholar 

  44. Fuziwara CS, Kimura ET. Insights into regulation of the miR-17-92 cluster of miRNAs in cancer. Front Med. 2015;2:64.

    Article  Google Scholar 

  45. Park SM, Gaur AB, Lengyel E, Peter ME. The miR-200 family determines the epithelial phenotype of cancer cells by targeting the E-cadherin repressors ZEB1 and ZEB2. Genes Dev. 2008;22(7):894–907.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Hussen BM, Salihi A, Abdullah ST, Rasul MF, Hidayat HJ, Hajiesmaeili M, et al. Signaling pathways modulated by miRNAs in breast cancer angiogenesis and new therapeutics. Pathol Res Pract. 2022;230:153764.

    Article  CAS  PubMed  Google Scholar 

  47. Ye J, Guo R, Shi Y, Qi F, Guo C, Yang L. miR-155 regulated inflammation response by the SOCS1-STAT3-PDCD4 axis in atherogenesis. Mediators Inflamm. 2016;2016:8060182.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Alexandrov P, Zhai Y, Li W, Lukiw W. Lipopolysaccharide-stimulated, NF-κB-, miRNA-146a- and miRNA-155-mediated molecular-genetic communication between the human gastrointestinal tract microbiome and the brain. Folia Neuropathol. 2019;57(3):211–9.

    Article  PubMed  Google Scholar 

  49. Kalsotra A, Wang K, Li PF, Cooper TA. MicroRNAs coordinate an alternative splicing network during mouse postnatal heart development. Genes Dev. 2010;24(7):653–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Suzuki H, Maruyama R, Yamamoto E, Kai M. Epigenetic alteration and microRNA dysregulation in cancer. Front Genet. 2013;4:258.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Dittmar RL, Sen S. Chapter 4 - microRNAs in exosomes in cancer. In: Chakrabarti DJ, Mitra DS, editors. Cancer and noncoding RNAs. Boston: Academic Press; 2018. p. 59–78.

    Chapter  Google Scholar 

  52. Menon A, Abd-Aziz N, Khalid K, Poh CL, Naidu R. miRNA: a promising therapeutic target in cancer. Int J Mol Sci. 2022;23(19):11502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Katti A, Diaz BJ, Caragine CM, Sanjana NE, Dow LE. CRISPR in cancer biology and therapy. Nat Rev Cancer. 2022;22(5):259–79.

    Article  CAS  PubMed  Google Scholar 

  54. Nguyen DD, Chang S. Development of novel therapeutic agents by inhibition of oncogenic MicroRNAs. Int J Mol Sci. 2017;19(1):65.

    Article  PubMed  PubMed Central  Google Scholar 

  55. He L, He X, Lim LP, De Stanchina E, Xuan Z, Liang Y, et al. A microRNA component of the p53 tumour suppressor network. Nature. 2007;447(7148):1130–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Wen D, Danquah M, Chaudhary AK, Mahato RI. Small molecules targeting microRNA for cancer therapy: promises and obstacles. J Control Release. 2015;219:237–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Trang P, Medina PP, Wiggins JF, Ruffino L, Kelnar K, Omotola M, et al. Regression of murine lung tumors by the let-7 microRNA. Oncogene. 2010;29(11):1580–7.

    Article  CAS  PubMed  Google Scholar 

  58. Daige CL, Wiggins JF, Priddy L, Nelligan-Davis T, Zhao J, Brown D. Systemic delivery of a miR34a mimic as a potential therapeutic for liver cancer. Mol Cancer Ther. 2014;13(10):2352–60.

    Article  CAS  PubMed  Google Scholar 

  59. Bader AG, Brown D, Winkler M. The promise of microRNA replacement therapy. Cancer Res. 2010;70(18):7027–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Zhang S, Cheng Z, Wang Y, Han T. The risks of miRNA therapeutics: in a drug target perspective. Drug Des Devel Ther. 2021;15:721–33.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Jin HY, Gonzalez-Martin A, Miletic AV, Lai M, Knight S, Sabouri-Ghomi M, et al. Transfection of microRNA mimics should be used with caution. Front Genet. 2015;6:340.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Søkilde R, Newie I, Persson H, Borg Å, Rovira C. Passenger strand loading in overexpression experiments using microRNA mimics. RNA Biol. 2015;12(8):787–91.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Ma L, Reinhardt F, Pan E, Soutschek J, Bhat B, Marcusson EG, et al. Therapeutic silencing of miR-10b inhibits metastasis in a mouse mammary tumor model. Nat Biotechnol. 2010;28(4):341–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Gong H, Chen H, Xiao P, Huang N, Han X, Zhang J, et al. miR-146a impedes the anti-aging effect of AMPK via NAMPT suppression and NAD+/SIRT inactivation. Signal Transduct Target Ther. 2022;7(1):66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Xu H, Guo S, Li W, Yu P. The circular RNA Cdr1as, via miR-7 and its targets, regulates insulin transcription and secretion in islet cells. Sci Rep. 2015;5:12453.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Su Y, Hou W, Zhang C, Ji P, Hu R, Zhang Q, et al. Long non-coding RNA ZFAS1 regulates cell proliferation and invasion in cervical cancer via the miR-190a-3p/KLF6 axis. Bioengineered. 2022;13(2):3840–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Ebert MS, Sharp PA. MicroRNA sponges: progress and possibilities. RNA. 2010;16(11):2043–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Ahmadi S, Sharifi M, Salehi R. Locked nucleic acid inhibits miR-92a-3p in human colorectal cancer, induces apoptosis and inhibits cell proliferation. Cancer Gene Ther. 2016;23(7):199–205.

    Article  CAS  PubMed  Google Scholar 

  69. De Cola A, Lamolinara A, Lanuti P, Rossi C, Iezzi M, Marchisio M, et al. MiR-205-5p inhibition by locked nucleic acids impairs metastatic potential of breast cancer cells. Cell Death Dis. 2018;9(8):821.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Ishida M, Selaru FM. miRNA-based therapeutic strategies. Curr Anesthesiol Rep. 2013;1(1):63–70.

    PubMed  Google Scholar 

  71. Owczarzy R, You Y, Groth CL, Tataurov AV. Stability and mismatch discrimination of locked nucleic acid-DNA duplexes. Biochemistry. 2011;50(43):9352–67.

    Article  CAS  PubMed  Google Scholar 

  72. Bader AG, Brown D, Stoudemire J, Lammers P. Developing therapeutic microRNAs for cancer. Gene Ther. 2011;18(12):1121–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Singh S, Narang AS, Mahato RI. Subcellular fate and off-target effects of siRNA, shRNA, and miRNA. Pharm Res. 2011;28(12):2996–3015.

    Article  CAS  PubMed  Google Scholar 

  74. Takagi S, Nakajima M, Kida K, Yamaura Y, Fukami T, Yokoi T. MicroRNAs regulate human hepatocyte nuclear factor 4alpha, modulating the expression of metabolic enzymes and cell cycle. J Biol Chem. 2010;285(7):4415–22.

    Article  CAS  PubMed  Google Scholar 

  75. Hong DS, Kang YK, Borad M, Sachdev J, Ejadi S, Lim HY, et al. Phase 1 study of MRX34, a liposomal miR-34a mimic, in patients with advanced solid tumours. Br J Cancer. 2020;122(11):1630–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Fortunato O, Iorio MV. The therapeutic potential of microRNAs in cancer: illusion or opportunity?. Pharmaceuticals. 2020;13(12):438.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. 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.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Huo W, Zhao G, Yin J, Ouyang X, Wang Y, Yang C, et al. Lentiviral CRISPR/Cas9 vector mediated miR-21 gene editing inhibits the epithelial to mesenchymal transition in ovarian cancer cells. J Cancer. 2017;8(1):57–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. El Fatimy R, Subramanian S, Uhlmann EJ, Krichevsky AM. Genome editing reveals glioblastoma addiction to microRNA-10b. Mol Ther. 2017;25(2):368–78.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Cyranoski D. CRISPR gene-editing tested in a person for the first time. Nature. 2016;539(7630):479.

    Article  CAS  PubMed  Google Scholar 

  81. Price C, Chen J. MicroRNAs in cancer biology and therapy: current status and perspectives. Genes Dis. 2014;1(1):53–63.

    Article  PubMed  PubMed Central  Google Scholar 

  82. Jing W, Zhang X, Sun W, Hou X, Yao Z, Zhu Y. CRISPR/CAS9-mediated genome editing of miRNA-155 inhibits proinflammatory cytokine production by RAW264.7 cells. Biomed Res Int. 2015;2015:326042.

    Article  PubMed  PubMed Central  Google Scholar 

  83. Bala S, Marcos M, Kodys K, Csak T, Catalano D, Mandrekar P, et al. Up-regulation of microRNA-155 in macrophages contributes to increased tumor necrosis factor α (TNFα) production via increased mRNA half-life in alcoholic liver disease. J Biol Chem. 2011;286(2):1436–44.

    Article  CAS  PubMed  Google Scholar 

  84. Li Z, Zhou X, Wei M, Gao X, Zhao L, Shi R, et al. In vitro and in vivo RNA inhibition by CD9-HuR functionalized exosomes encapsulated with miRNA or CRISPR/dCas9. Nano Lett. 2018;19(1):19–28.

    Article  CAS  PubMed  Google Scholar 

  85. Narayanan A, Hill-Teran G, Moro A, Ristori E, Kasper DM, Roden A, C, et al. In vivo mutagenesis of miRNA gene families using a scalable multiplexed CRISPR/Cas9 nuclease system. Sci Rep. 2016;6:32386.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Zhang YH, Wu LZ, Liang HL, Yang Y, Qiu J, Kan Q, et al. Pulmonary surfactant synthesis in miRNA-26a-1/miRNA-26a-2 double knockout mice generated using the CRISPR/Cas9 system. Am J Transl Res. 2017;9(2):355–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Godden AM, Antonaci M, Ward NJ, Van Der Lee M, Abu-Daya A, Guille M, et al. An efficient miRNA knockout approach using CRISPR-Cas9 in Xenopus. Dev Biol. 2022;483:66–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Senís E, Mockenhaupt S, Rupp D, Bauer T, Paramasivam N, Knapp B, et al. TALEN/CRISPR-mediated engineering of a promoterless anti-viral RNAi hairpin into an endogenous miRNA locus. Nucl Acids Res. 2017;45(1):e3.

    Article  PubMed  Google Scholar 

  89. Wang Z, Sun X, Zhang X, Dong B, Yu H. Development of a miRNA sensor by an inducible CRISPR-Cas9 construct in Ciona embryogenesis. Mol Biotechnol. 2021;63(7):613–20.

    Article  CAS  PubMed  Google Scholar 

  90. Zhang Z, Ursin R, Mahapatra S, Gallicano GI. CRISPR/CAS9 ablation of individual miRNAs from a miRNA family reveals their individual efficacies for regulating cardiac differentiation. Mech Dev. 2018;150:10–20.

    Article  CAS  PubMed  Google Scholar 

  91. Michaels YS, Wu Q, Fulga TA. Interrogation of functional miRNA–target interactions by CRISPR/Cas9 genome engineering. In: Dalmay T, editor. MicroRNA detection and target identification. Humana, New York, NY: Springer; 2017. p. 79–97.

    Chapter  Google Scholar 

  92. Matano M, Date S, Shimokawa M, Takano A, Fujii M, Ohta Y, et al. Modeling colorectal cancer using CRISPR-Cas9–mediated engineering of human intestinal organoids. Nat Med. 2015;21(3):256–62.

    Article  CAS  PubMed  Google Scholar 

  93. Zhao Y, Shi C, Zhao Y, Xin Z, Liu P, Zhang C, et al. Construction of miRNA-29b1 knockout mice based on CRISPR/Cas9 technology. Chin J Comp Med. 2016;26(12):1–4.

    Google Scholar 

  94. Kurata JS, Lin RJ. MicroRNA-focused CRISPR-Cas9 library screen reveals fitness-associated miRNAs. RNA. 2018;24(7):966–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Lataniotis L, Albrecht A, Kok FO, Monfries CA, Benedetti L, Lawson ND, et al. CRISPR/Cas9 editing reveals novel mechanisms of clustered microRNA regulation and function. Sci Rep. 2017;7(1):8585.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Voytas DF. Plant genome engineering with sequence-specific nucleases. Annu Rev Plant Biol. 2013;64:327–50.

    Article  CAS  PubMed  Google Scholar 

  97. Zhang JP, Li XL, Li GH, Chen W, Arakaki C, Botimer GD, et al. Efficient precise knockin with a double cut HDR donor after CRISPR/Cas9-mediated double-stranded DNA cleavage. Genome Biol. 2017;18(1):35.

    Article  PubMed  PubMed Central  Google Scholar 

  98. Sander JD, Dahlborg EJ, Goodwin MJ, Cade L, Zhang F, Cifuentes D, et al. Selection-free zinc-finger-nuclease engineering by context-dependent assembly (CoDA). Nat Methods. 2011;8(1):67–9.

    Article  CAS  PubMed  Google Scholar 

  99. Cermak T, Doyle EL, Christian M, Wang L, Zhang Y, Schmidt C, et al. Efficient design and assembly of custom TALEN and other TAL effector-based constructs for DNA targeting. Nucleic Acids Res. 2011;39(12):e82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Bin Moon S, Lee JM, Kang JG, Lee N-E, Ha D-I, Kim DY, et al. Highly efficient genome editing by CRISPR-Cpf1 using CRISPR RNA with a uridinylate-rich 3′-overhang. Nat Commun. 2018;9(1):3651.

    Article  PubMed  PubMed Central  Google Scholar 

  101. Wang JY, Pausch P, Doudna JA. Structural biology of CRISPR–Cas immunity and genome editing enzymes. Nat Rev Microbiol. 2022;20(11):641–56.

    Article  CAS  PubMed  Google Scholar 

  102. Zhang H, Qin C, An C, Zheng X, Wen S, Chen W, et al. Application of the CRISPR/Cas9-based gene editing technique in basic research, diagnosis, and therapy of cancer. Mol Cancer. 2021;20(1):126.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Peng S, Tan Z, Chen S, Lei C, Nie Z. Integrating CRISPR-Cas12a with a DNA circuit as a generic sensing platform for amplified detection of microRNA. Chem Sci. 2020;11(28):7362–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Anzalone AV, Koblan LW, Liu DR. Genome editing with CRISPR-Cas nucleases, base editors, transposases and prime editors. Nat Biotechnol. 2020;38(7):824–44.

    Article  CAS  PubMed  Google Scholar 

  105. Gleditzsch D, Pausch P, Müller-Esparza H, Özcan A, Guo X, Bange G, et al. PAM identification by CRISPR-Cas effector complexes: diversified mechanisms and structures. RNA Biol. 2019;16(4):504–17.

    Article  PubMed  Google Scholar 

  106. Makarova KS, Wolf YI, Iranzo J, Shmakov SA, Alkhnbashi OS, Brouns SJJ, et al. Evolutionary classification of CRISPR-Cas systems: a burst of class 2 and derived variants. Nat Rev Microbiol. 2020;18(2):67–83.

    Article  CAS  PubMed  Google Scholar 

  107. Pelea O, Fulga TA, Sauka-Spengler T. RNA-responsive gRNAs for controlling CRISPR activity: current advances, future directions, and potential applications. CRISPR J. 2022;5(5):642–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Nishimasu H, Ran FA, Hsu PD, Konermann S, Shehata SI, Dohmae N, et al. Crystal structure of Cas9 in complex with guide RNA and target DNA. Cell. 2014;156(5):935–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Zhao Z, Li C, Tong F, Deng J, Huang G, Sang Y. Review of applications of CRISPR-Cas9 gene-editing technology in cancer research. Biol Proced Online. 2021;23(1):14.

    Article  PubMed  PubMed Central  Google Scholar 

  110. Jiang F, Zhou K, Ma L, Gressel S, Doudna JA. Structural biology. A Cas9-guide RNA complex preorganized for target DNA recognition. Science. 2015;348(6242):1477–81.

    Article  CAS  PubMed  Google Scholar 

  111. Anders C, Niewoehner O, Duerst A, Jinek M. Structural basis of PAM-dependent target DNA recognition by the Cas9 endonuclease. Nature. 2014;513(7519):569–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Pacesa M, Loeff L, Querques I, Muckenfuss LM, Sawicka M, Jinek M. R-loop formation and conformational activation mechanisms of Cas9. Nature. 2022;609(7925):191–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Globyte V, Lee SH, Bae T, Kim JS, Joo C. CRISPR/Cas9 searches for a protospacer adjacent motif by lateral diffusion. Embo J. 2019;38(4):e99466.

    Article  PubMed  Google Scholar 

  114. Cofsky JC, Soczek KM, Knott GJ, Nogales E, Doudna JA. CRISPR-Cas9 bends and twists DNA to read its sequence. Nat Struct Mol Biol. 2022;29(4):395–402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Jiang F, Taylor DW, Chen JS, Kornfeld JE, Zhou K, Thompson AJ, et al. Structures of a CRISPR-Cas9 R-loop complex primed for DNA cleavage. Science. 2016;351(6275):867–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Ivanov IE, Wright AV, Cofsky JC, Aris KDP, Doudna JA, Bryant Z. Cas9 interrogates DNA in discrete steps modulated by mismatches and supercoiling. Proc Natl Acad Sci U S A. 2020;117(11):5853–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Cammaerts S, Strazisar M, De Rijk P, Del Favero J. Genetic variants in microRNA genes: impact on microRNA expression, function, and disease. Front Genet. 2015;6:186.

    Article  PubMed  PubMed Central  Google Scholar 

  118. Zhou J, Deng K, Cheng Y, Zhong Z, Tian L, Tang X, et al. CRISPR-Cas9 based genome editing reveals new insights into microRNA function and regulation in rice. Front Plant Sci. 2017;8:1598.

    Article  PubMed  PubMed Central  Google Scholar 

  119. Zhao Y, Dai Z, Liang Y, Yin M, Ma K, He M, et al. Sequence-specific inhibition of microRNA via CRISPR/CRISPRi system. Sci Rep. 2014;4:3943.

    Article  PubMed  PubMed Central  Google Scholar 

  120. Sheng P, Fields C, Aadland K, Wei T, Kolaczkowski O, Gu T, et al. Dicer cleaves 5’-extended microRNA precursors originating from RNA polymerase II transcription start sites. Nucleic Acids Res. 2018;46(11):5737–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Abdollah NA, Theva Das K, Narazah M, Abdul Razak SR. Sequence-specific inhibition of microRNA-130a gene by CRISPR/Cas9 system in breast cancer cell line. J Phys Conf Ser. 2017;851:012037. https://doi.org/10.1088/1742-6596/851/1/012037.

    Article  CAS  Google Scholar 

  122. Park HH, Triboulet R, Bentler M, Guda S, Du P, Xu H, et al. DROSHA knockout leads to enhancement of viral titers for vectors encoding miRNA-adapted shRNAs. Mol Ther Nucleic Acids. 2018;12:591–9.

    Article  PubMed  PubMed Central  Google Scholar 

  123. Shmakov S, Smargon A, Scott D, Cox D, Pyzocha N, Yan W, et al. Diversity and evolution of class 2 CRISPR-Cas systems. Nat Rev Microbiol. 2017;15(3):169–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Carabias A, Fuglsang A, Temperini P, Pape T, Sofos N, Stella S, et al. Structure of the mini-RNA-guided endonuclease CRISPR-Cas12j3. Nat Commun. 2021;12(1):4476.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Tsuchida CA, Zhang S, Doost MS, Zhao Y, Wang J, et al. Chimeric CRISPR-CasX enzymes and guide RNAs for improved genome editing activity. Mol Cell. 2022;82(6):1199–209.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Huang CJ, Adler BA, Doudna JA. A naturally DNase-free CRISPR-Cas12c enzyme silences gene expression. Mol Cell. 2022;82(11):2148-60.e4.

    Article  CAS  PubMed  Google Scholar 

  127. Xiao R, Li Z, Wang S, Han R, Chang L. Structural basis for substrate recognition and cleavage by the dimerization-dependent CRISPR-Cas12f nuclease. Nucleic Acids Res. 2021;49(7):4120–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Takeda SN, Nakagawa R, Okazaki S, Hirano H, Kobayashi K, Kusakizako T, et al. Structure of the miniature type V-F CRISPR-Cas effector enzyme. Mol Cell. 2021;81(3):558-70.e3.

    Article  CAS  PubMed  Google Scholar 

  129. Pausch P, Soczek KM, Herbst DA, Tsuchida CA, Al-Shayeb B, Banfield JF, et al. DNA interference states of the hypercompact CRISPR–CasΦ effector. Nat Struct Mol Biol. 2021;28(8):652–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Li Z, Zhang H, Xiao R, Han R, Chang L. Cryo-EM structure of the RNA-guided ribonuclease Cas12g. Nat Chem Biol. 2021;17(4):387–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Dong D, Ren K, Qiu X, Zheng J, Guo M, Guan X, et al. The crystal structure of Cpf1 in complex with CRISPR RNA. Nature. 2016;532(7600):522–6.

    Article  CAS  PubMed  Google Scholar 

  132. Swarts DC, Van Der Oost J, Jinek M. Structural basis for guide RNA processing and seed-dependent DNA targeting by CRISPR-Cas12a. Mol Cell. 2017;66(2):221-33.e4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Stella S, Alcón P, Montoya G. Structure of the Cpf1 endonuclease R-loop complex after target DNA cleavage. Nature. 2017;546(7659):559–63.

    Article  CAS  PubMed  Google Scholar 

  134. Stella S, Mesa P, Thomsen J, Paul B, Alcón P, Jensen SB, et al. Conformational activation promotes CRISPR-Cas12a catalysis and resetting of the endonuclease activity. Cell. 2018;175(7):1856-71.e21.

    Article  CAS  PubMed  Google Scholar 

  135. Zetsche B, Gootenberg JS, Abudayyeh OO, Slaymaker IM, Makarova KS, Essletzbichler P, et al. Cpf1 is a single RNA-guided endonuclease of a class 2 CRISPR-Cas system. Cell. 2015;163(3):759–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Huang X, Sun W, Cheng Z, Chen M, Li X, Wang J, et al. Structural basis for two metal-ion catalysis of DNA cleavage by Cas12i2. Nat Commun. 2020;11(1):5241.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Swarts DC, Jinek M. Mechanistic insights into the cis- and trans-acting DNase activities of Cas12a. Mol Cell. 2019;73(3):589-600.e4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Jiang W, Singh J, Allen A, Li Y, Kathiresan V, Qureshi O, et al. CRISPR-Cas12a nucleases bind flexible dna duplexes without RNA/DNA complementarity. ACS Omega. 2019;4(17):17140–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Paul B, Chaubet L, Verver DE, Montoya G. Mechanics of CRISPR-Cas12a and engineered variants on λ-DNA. Nucleic Acids Res. 2022;50(9):5208–25.

    Article  CAS  PubMed  Google Scholar 

  140. Jiang Q, Meng X, Meng L, Chang N, Xiong J, Cao H, et al. Small indels induced by CRISPR/Cas9 in the 5’ region of microRNA lead to its depletion and Drosha processing retardance. RNA Biol. 2014;11(10):1243–9.

    Article  PubMed  Google Scholar 

  141. Chen X. MicroRNA biogenesis and function in plants. FEBS Lett. 2005;579(26):5923–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Budak H, Akpinar BA. Plant miRNAs: biogenesis, organization and origins. Funct Integr Genomics. 2015;15(5):523–31.

    Article  CAS  PubMed  Google Scholar 

  143. Ha M, Kim VN. Regulation of microRNA biogenesis. Nat Rev Mol Cell Biol. 2014;15(8):509–24.

    Article  CAS  PubMed  Google Scholar 

  144. Lewis B, Shih IH, Jones-Rhoades M, Bartel D, Burge C. Prediction of mammalian MicroRNA Targets. Cell. 2004;115:787–98.

    Article  Google Scholar 

  145. Bartel DP. MicroRNAs: target recognition and regulatory functions. Cell. 2009;136(2):215–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Chung PJ, Chung H, Oh N, Choi J, Bang SW, Jung SE, et al. Efficiency of recombinant CRISPR/rCas9-mediated miRNA gene editing in rice. Int J Mol Sci. 2020;21(24):9606.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Burke JM, Kelenis DP, Kincaid RP, Sullivan CS. A central role for the primary microRNA stem in guiding the position and efficiency of Drosha processing of a viral pri-miRNA. RNA. 2014;20(7):1068–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Friedland AE, Tzur YB, Esvelt KM, Colaiácovo MP, Church GM, Calarco JA. Heritable genome editing in C. elegans via a CRISPR-Cas9 system. Nat Methods. 2013;10(8):741–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Wu X, Kriz AJ, Sharp PA. Target specificity of the CRISPR-Cas9 system. Quant Biol. 2014;2(2):59–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Cao J, Xiao Q, Yan Q. The multiplexed CRISPR targeting platforms. Drug Discov Today Technol. 2018;28:53–61.

    Article  PubMed  PubMed Central  Google Scholar 

  151. Ernst MPT, Broeders M, Herrero-Hernandez P, Oussoren E, Van Der Ploeg AT, Pijnappel W. Ready for repair? Gene editing enters the clinic for the treatment of human disease. Mol Ther Methods Clin Dev. 2020;18:532–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Cullot G, Boutin J, Toutain J, Prat F, Pennamen P, Rooryck C, et al. CRISPR-Cas9 genome editing induces megabase-scale chromosomal truncations. Nat Commun. 2019;10(1):1136.

    Article  PubMed  PubMed Central  Google Scholar 

  153. Kosicki M, Tomberg K, Bradley A. Repair of double-strand breaks induced by CRISPR-Cas9 leads to large deletions and complex rearrangements. Nat Biotechnol. 2018;36(8):765–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Karvelis T, Bigelyte G, Young JK, Hou Z, Zedaveinyte R, Budre K, et al. PAM recognition by miniature CRISPR-Cas12f nucleases triggers programmable double-stranded DNA target cleavage. Nucleic Acids Res. 2020;48(9):5016–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Pausch P, Al-Shayeb B, Bisom-Rapp E, Tsuchida CA, Li Z, Cress BF, et al. CRISPR-CasΦ from huge phages is a hypercompact genome editor. Science. 2020;369(6501):333–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Harrington LB, Burstein D, Chen JS, Paez-Espino D, Ma E, Witte IP, et al. Programmed DNA destruction by miniature CRISPR-Cas14 enzymes. Science. 2018;362(6416):839–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Strohkendl I, Saifuddin FA, Rybarski JR, Finkelstein IJ, Russell R. kinetic basis for DNA target specificity of CRISPR-Cas12a. Mol Cell. 2018;71(5):816-24.e3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Kleinstiver BP, Tsai SQ, Prew MS, Nguyen NT, Welch MM, Lopez JM, et al. Genome-wide specificities of CRISPR-Cas Cpf1 nucleases in human cells. Nat Biotechnol. 2016;34(8):869–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Kim D, Kim J, Hur JK, Been KW, Yoon SH, Kim JS. Genome-wide analysis reveals specificities of Cpf1 endonucleases in human cells. Nat Biotechnol. 2016;34(8):863–8.

    Article  CAS  PubMed  Google Scholar 

  160. Zetsche B, Heidenreich M, Mohanraju P, Fedorova I, Kneppers J, Degennaro EM, et al. Multiplex gene editing by CRISPR-Cpf1 using a single crRNA array. Nat Biotechnol. 2017;35(1):31–4.

    Article  CAS  PubMed  Google Scholar 

  161. Fonfara I, Richter H, Bratovič M, Le Rhun A, Charpentier E. The CRISPR-associated DNA-cleaving enzyme Cpf1 also processes precursor CRISPR RNA. Nature. 2016;532(7600):517–21.

    Article  CAS  PubMed  Google Scholar 

  162. Kim SK, Kim H, Ahn WC, Park KH, Woo EJ, Lee DH, et al. Efficient transcriptional gene repression by type V-A CRISPR-Cpf1 from eubacterium eligens. ACS Synth Biol. 2017;6(7):1273–82.

    Article  CAS  PubMed  Google Scholar 

  163. Zhang X, Wang J, Cheng Q, Zheng X, Zhao G, Wang J. Multiplex gene regulation by CRISPR-ddCpf1. Cell Discov. 2017;3:17018.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Cong L, Ran FA, Cox D, Lin S, Barretto R, Habib N, et al. Multiplex genome engineering using CRISPR/Cas systems. Science. 2013;339(6121):819–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Hu JH, Miller SM, Geurts MH, Tang W, Chen L, Sun N, et al. Evolved Cas9 variants with broad PAM compatibility and high DNA specificity. Nature. 2018;556(7699):57–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Bi H, Fei Q, Li R, Liu B, Xia R, Char SN, et al. Disruption of miRNA sequences by TALENs and CRISPR/Cas9 induces varied lengths of miRNA production. Plant Biotechnol J. 2020;18(7):1526–36.

    Article  CAS  PubMed  Google Scholar 

  167. Paul B, Montoya G. CRISPR-Cas12a: functional overview and applications. Biomed J. 2020;43(1):8–17.

    Article  PubMed  PubMed Central  Google Scholar 

  168. Abels ER, Maas SLN, Nieland L, Wei Z, Cheah PS, Tai E, et al. Glioblastoma-associated microglia reprogramming is mediated by functional transfer of extracellular miR-21. Cell Rep. 2019;28(12):3105-19.e7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Van Der Vos KE, Abels ER, Zhang X, Lai C, Carrizosa E, Oakley D, et al. Directly visualized glioblastoma-derived extracellular vesicles transfer RNA to microglia/macrophages in the brain. Neuro Oncol. 2016;18(1):58–69.

    Article  PubMed  Google Scholar 

  170. Nieland L, Van Solinge TS, Cheah PS, Morsett LM, El Khoury J, Rissman JI, et al. CRISPR-Cas knockout of miR21 reduces glioma growth. Mol Ther Oncolytics. 2022;25:121–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Gaudelli NM, Komor AC, Rees HA, Packer MS, Badran AH, Bryson DI, et al. Programmable base editing of A•T to G•C in genomic DNA without DNA cleavage. Nature. 2017;551(7681):464–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Xie H, Tang L, He X, Liu X, Zhou C, Liu J, et al. SaCas9 requires 5’-NNGRRT-3’ PAM for sufficient cleavage and possesses higher cleavage activity than SpCas9 or FnCpf1 in human cells. Biotechnol J. 2018;13(4):e1700561.

    Article  PubMed  Google Scholar 

  173. Kleinstiver BP, Prew MS, Tsai SQ, Nguyen NT, Topkar VV, Zheng Z, et al. Broadening the targeting range of Staphylococcus aureus CRISPR-Cas9 by modifying PAM recognition. Nat Biotechnol. 2015;33(12):1293–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Pickar-Oliver A, Gersbach CA. The next generation of CRISPR-Cas technologies and applications. Nat Rev Mol Cell Biol. 2019;20(8):490–507.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Chatterjee P, Jakimo N, Jacobson JM. Minimal PAM specificity of a highly similar SpCas9 ortholog. Sci Adv. 2018;4(10):eaau0766.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Yamada M, Watanabe Y, Gootenberg JS, Hirano H, Ran FA, Nakane T, et al. Crystal structure of the minimal Cas9 from Campylobacter jejuni reveals the molecular diversity in the CRISPR-Cas9 systems. Mol Cell. 2017;65(6):1109-21.e3.

    Article  CAS  PubMed  Google Scholar 

  177. Schindele P, Puchta H. Engineering CRISPR/LbCas12a for highly efficient, temperature-tolerant plant gene editing. Plant Biotechnol J. 2020;18(5):1118.

    Article  PubMed  Google Scholar 

  178. Kleinstiver BP, Sousa AA, Walton RT, Tak YE, Hsu JY, Clement K, et al. Engineered CRISPR-Cas12a variants with increased activities and improved targeting ranges for gene, epigenetic and base editing. Nat Biotechnol. 2019;37(3):276–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Morisaka H, Yoshimi K, Okuzaki Y, Gee P, Kunihiro Y, Sonpho E, et al. CRISPR-Cas3 induces broad and unidirectional genome editing in human cells. Nat Commun. 2019;10(1):5302.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Hsu JY, Grünewald J, Szalay R, Shih J, Anzalone AV, Lam KC, et al. PrimeDesign software for rapid and simplified design of prime editing guide RNAs. Nat Commun. 2021;12(1):1034.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Kim HK, Lee S, Kim Y, Park J, Min S, Choi JW, et al. High-throughput analysis of the activities of xCas9, SpCas9-NG and SpCas9 at matched and mismatched target sequences in human cells. Nat Biomed Eng. 2020;4(1):111–24.

    Article  CAS  PubMed  Google Scholar 

  182. Kang SH, Lee WJ, An JH, Lee JH, Kim YH, Kim H, et al. Prediction-based highly sensitive CRISPR off-target validation using target-specific DNA enrichment. Nat Commun. 2020;11(1):3596.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Zhang XH, Tee LY, Wang XG, Huang QS, Yang SH. Off-target effects in CRISPR/Cas9-mediated genome engineering. Mol Ther Nucleic Acids. 2015;4:e264.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Ren X, Yang Z, Xu J, Sun J, Mao D, Hu Y, et al. Enhanced specificity and efficiency of the CRISPR/Cas9 system with optimized sgRNA parameters in drosophila. Cell Rep. 2014;9(3):1151–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Matson AW, Hosny N, Swanson ZA, Hering BJ, Burlak C. Optimizing sgRNA length to improve target specificity and efficiency for the GGTA1 gene using the CRISPR/Cas9 gene editing system. PLoS One. 2019;14(12):e0226107.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Gabriel R, Von Kalle C, Schmidt M. Mapping the precision of genome editing. Nat Biotechnol. 2015;33(2):150–2.

    Article  CAS  PubMed  Google Scholar 

  187. Clement K, Rees H, Canver MC, Gehrke JM, Farouni R, Hsu JY, et al. CRISPResso2 provides accurate and rapid genome editing sequence analysis. Nat Biotechnol. 2019;37(3):224–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Güell M, Yang L, Church GM. Genome editing assessment using CRISPR genome analyzer (CRISPR-GA). Bioinformatics. 2014;30(20):2968–70.

    Article  PubMed  PubMed Central  Google Scholar 

  189. Concordet J-P, Haeussler M. CRISPOR: intuitive guide selection for CRISPR/Cas9 genome editing experiments and screens. Nucleic Acids Res. 2018;46(W1):W242–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Liu H, Wei Z, Dominguez A, Li Y, Wang X, Qi LS. CRISPR-ERA: a comprehensive design tool for CRISPR-mediated gene editing, repression and activation. Bioinformatics. 2015;31(22):3676–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Montague TG, Cruz JM, Gagnon JA, Church GM, Valen E. CHOPCHOP: a CRISPR/Cas9 and TALEN web tool for genome editing. Nucleic Acids Res. 2014;42(W1):W401–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Park J, Bae S, Kim JS. Cas-Designer: a web-based tool for choice of CRISPR-Cas9 target sites. Bioinformatics. 2015;31(24):4014–6.

    Article  CAS  PubMed  Google Scholar 

  193. Meier JA, Zhang F, Sanjana NE. GUIDES: sgRNA design for loss-of-function screens. Nat Methods. 2017;14(9):831–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Moreno-Mateos MA, Vejnar CE, Beaudoin J-D, Fernandez JP, Mis EK, Khokha MK, et al. CRISPRscan: designing highly efficient sgRNAs for CRISPR-Cas9 targeting in vivo. Nat Methods. 2015;12(10):982–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Stemmer M, Thumberger T, Del Sol KM, Wittbrodt J, Mateo JL. CCTop: an intuitive, flexible and reliable CRISPR/Cas9 target prediction tool. PLoS One. 2015;10(4):e0124633.

    Article  PubMed  PubMed Central  Google Scholar 

  196. Abadi S, Yan WX, Amar D, Mayrose I. A machine learning approach for predicting CRISPR-Cas9 cleavage efficiencies and patterns underlying its mechanism of action. PLoS Comput Biol. 2017;13(10):e1005807.

    Article  PubMed  PubMed Central  Google Scholar 

  197. Chuai G, Ma H, Yan J, Chen M, Hong N, Xue D, et al. DeepCRISPR: optimized CRISPR guide RNA design by deep learning. Genome Biol. 2018;19(1):80.

    Article  PubMed  PubMed Central  Google Scholar 

  198. Wong N, Liu W, Wang X. WU-CRISPR: characteristics of functional guide RNAs for the CRISPR/Cas9 system. Genome Biol. 2015;16(1):218.

    Article  PubMed  PubMed Central  Google Scholar 

  199. Varshney GK, Zhang S, Pei W, Adomako-Ankomah A, Fohtung J, Schaffer K, et al. CRISPRz: a database of zebrafish validated sgRNAs. Nucleic Acids Res. 2016;44(D1):D822–6.

    Article  CAS  PubMed  Google Scholar 

  200. Zhao G, Li J, Tang Y. AsCRISPR: a web server for allele-specific sgRNA design in precision medicine. CRISPR J. 2020;3(6):512–22.

    Article  CAS  PubMed  Google Scholar 

  201. Chen W, Zhang G, Li J, Zhang X, Huang S, Xiang S, et al. CRISPRlnc: a manually curated database of validated sgRNAs for lncRNAs. Nucleic Acids Res. 2019;47(D1):D63–8.

    Article  CAS  PubMed  Google Scholar 

  202. Pattanayak V, Lin S, Guilinger JP, Ma E, Doudna JA, Liu DR. High-throughput profiling of off-target DNA cleavage reveals RNA-programmed Cas9 nuclease specificity. Nat Biotechnol. 2013;31(9):839–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Tyumentseva MA, Tyumentsev AI, Akimkin VG. Protocol for assessment of the efficiency of CRISPR/Cas RNP delivery to different types of target cells. PLoS One. 2021;16(11):e0259812.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Basila M, Kelley ML, Smith AVB. Minimal 2’-O-methyl phosphorothioate linkage modification pattern of synthetic guide RNAs for increased stability and efficient CRISPR-Cas9 gene editing avoiding cellular toxicity. PLoS One. 2017;12(11):e0188593.

    Article  PubMed  PubMed Central  Google Scholar 

  205. Rahdar M, Mcmahon MA, Prakash TP, Swayze EE, Bennett CF, Cleveland DW. Synthetic CRISPR RNA-Cas9-guided genome editing in human cells. Proc Natl Acad Sci U S A. 2015;112(51):E7110–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  206. Wang X, Wang Y, Wu X, Wang J, Wang Y, Qiu Z, et al. Unbiased detection of off-target cleavage by CRISPR-Cas9 and TALENs using integrase-defective lentiviral vectors. Nat Biotechnol. 2015;33(2):175–8.

    Article  CAS  PubMed  Google Scholar 

  207. Thomson DW, Bracken CP, Goodall GJ. Experimental strategies for microRNA target identification. Nucleic Acids Res. 2011;39(16):6845–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Wu S, Huang S, Ding J, Zhao Y, Liang L, Liu T, et al. Multiple microRNAs modulate p21Cip1/Waf1 expression by directly targeting its 3’ untranslated region. Oncogene. 2010;29(15):2302–8.

    Article  CAS  PubMed  Google Scholar 

  209. Kabadi AM, Ousterout DG, Hilton IB, Gersbach CA. Multiplex CRISPR/Cas9-based genome engineering from a single lentiviral vector. Nucleic Acids Res. 2014;42(19):e147.

    Article  PubMed  PubMed Central  Google Scholar 

  210. Kato Y, Tabata H, Sato K, Nakamura M, Saito I, Nakanishi T. Adenovirus vectors expressing eight multiplex guide RNAs of CRISPR/Cas9 efficiently disrupted diverse hepatitis B virus gene derived from heterogeneous patient. Int J Mol Sci. 2021;22(19):10570.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Asmamaw MM. Viral Vectors for the in Vivo Delivery of CRISPR components: advances and challenges. Front Bioeng Biotechnol. 2022;10:895713.

    Article  Google Scholar 

  212. Mout R, Ray M, Lee YW, Scaletti F, Rotello VM. In vivo delivery of CRISPR/Cas9 for therapeutic gene editing: progress and challenges. Bioconjug Chem. 2017;28(4):880–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Van Haasteren J, Li J, Scheideler OJ, Murthy N, Schaffer DV. The delivery challenge: fulfilling the promise of therapeutic genome editing. Nat Biotechnol. 2020;38(7):845–55.

    Article  PubMed  Google Scholar 

  214. Yip BH. Recent advances in CRISPR/Cas9 delivery strategies. Biomolecules. 2020;10(6):839.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Ricobaraza A, Gonzalez-Aparicio M, Mora-Jimenez L, Lumbreras S, Hernandez-Alcoceba R. High-capacity adenoviral vectors: expanding the scope of gene therapy. Int J Mol Sci. 2020;21(10):3643.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Ahi YS, Bangari DS, Mittal SK. Adenoviral vector immunity: its implications and circumvention strategies. Curr Gene Ther. 2011;11(4):307–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Luo Y, Xu X, An X, Sun X, Wang S, Zhu D. Targeted inhibition of the miR-199a/214 cluster by CRISPR interference augments the tumor tropism of human induced pluripotent stem cell-derived neural stem cells under hypoxic condition. Stem Cells Int. 2016;2016:3598542.

    Article  PubMed  PubMed Central  Google Scholar 

  218. Hu YC. Baculovirus as a highly efficient expression vector in insect and mammalian cells. Acta Pharmacol Sin. 2005;26(4):405–16.

    Article  CAS  PubMed  Google Scholar 

  219. Drobna-Śledzińska M, Maćkowska-Maślak N, Jaksik R, Dąbek P, Witt M, Dawidowska M. CRISPRi for specific inhibition of miRNA clusters and miRNAs with high sequence homology. Sci Rep. 2022;12(1):6297.

    Article  PubMed  PubMed Central  Google Scholar 

  220. Potter H, Heller R. Transfection by electroporation. Curr Protoc Mol Biol. 2018;121:9.3.1-13.

    Article  CAS  PubMed  Google Scholar 

  221. Lino CA, Harper JC, Carney JP, Timlin JA. Delivering CRISPR: a review of the challenges and approaches. Drug Deliv. 2018;25(1):1234–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Dasgupta I, Chatterjee A. Recent advances in miRNA delivery systems. Methods Protoc. 2021;4(1):10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Chen F, Alphonse M, Liu Q. Strategies for nonviral nanoparticle-based delivery of CRISPR/Cas9 therapeutics. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2020;12(3):e1609.

    Article  PubMed  Google Scholar 

  224. Ran F, Cong L, Yan WX, Scott DA, Gootenberg JS, Kriz AJ, et al. In vivo genome editing using Staphylococcus aureus Cas9. Nature. 2015;520(7546):186–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Aquino-Jarquin G. CRISPR-Cas14 is now part of the artillery for gene editing and molecular diagnostic. Nanomedicine. 2019;18:428–31.

    Article  CAS  PubMed  Google Scholar 

  226. Chew WL, Tabebordbar M, Cheng JK, Mali P, Wu EY, Ng AH, et al. A multifunctional AAV-CRISPR-Cas9 and its host response. Nat Methods. 2016;13(10):868–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Alton EW, Armstrong DK, Ashby D, Bayfield KJ, Bilton D, Bloomfield EV, et al. Repeated nebulisation of non-viral CFTR gene therapy in patients with cystic fibrosis: a randomised, double-blind, placebo-controlled, phase 2b trial. Lancet Respir Med. 2015;3(9):684–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  228. Behr M, Zhou J, Xu B, Zhang H. In vivo delivery of CRISPR-Cas9 therapeutics: progress and challenges. Acta Pharm Sin B. 2021;11(8):2150–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  229. Lin Y, Wu J, Gu W, Huang Y, Tong Z, Huang L, et al. Exosome-liposome hybrid nanoparticles deliver CRISPR/Cas9 system in MSCs. Adv Sci (Weinh). 2018;5(4):1700611.

    Article  PubMed  Google Scholar 

  230. Duan L, Ouyang K, Xu X, Xu L, Wen C, Zhou X, et al. Nanoparticle delivery of CRISPR/Cas9 for genome editing. Front Genet. 2021;12:673286.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  231. Steinmetz NF, Lim S, Sainsbury F. Protein cages and virus-like particles: from fundamental insight to biomimetic therapeutics. Biomater Sci. 2020;8(10):2771–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

MT, BMH, and WB designed and supervised the study. SGF and BMH wrote the biogenesis of miRNAs and the challenges of current miRNA-based cancer therapy. AB and MR wrote the innovative advances in CRISPR/Cas miRNA-editing technology. MFR, FAA, and MCG wrote the regions of miRNAs gene targeted with CRISPR/Cas and the advantages of miRNAs targeting. SRA and HJH collected the data and manuscript drafting. GSMF, FAA, and AS collected the data and designed the figures and tables. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Wojciech Branicki or Mohammad Taheri.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hussen, B.M., Rasul, M.F., Abdullah, S.R. et al. Targeting miRNA by CRISPR/Cas in cancer: advantages and challenges. Military Med Res 10, 32 (2023). https://doi.org/10.1186/s40779-023-00468-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40779-023-00468-6

Keywords