Skip to main content

CRISPR/Cas9 application in cancer therapy: a pioneering genome editing tool

Abstract

The progress of genetic engineering in the 1970s brought about a paradigm shift in genome editing technology. The clustered regularly interspaced short palindromic repeats/CRISPR associated protein 9 (CRISPR/Cas9) system is a flexible means to target and modify particular DNA sequences in the genome. Several applications of CRISPR/Cas9 are presently being studied in cancer biology and oncology to provide vigorous site-specific gene editing to enhance its biological and clinical uses. CRISPR's flexibility and ease of use have enabled the prompt achievement of almost any preferred alteration with greater efficiency and lower cost than preceding modalities. Also, CRISPR/Cas9 technology has recently been applied to improve the safety and efficacy of chimeric antigen receptor (CAR)-T cell therapies and defeat tumor cell resistance to conventional treatments such as chemotherapy and radiotherapy. The current review summarizes the application of CRISPR/Cas9 in cancer therapy. We also discuss the present obstacles and contemplate future possibilities in this context.

Introduction

Genome editing tools have offered great advantages to the biological sciences [1, 2]. Various techniques, including zinc finger endonuclease (ZFN), transcription activator-like effector nuclease (TALEN), and the clustered regularly interspaced short palindromic repeats/CRISPR associated nuclease (CRISPR/Cas) system, have now been developed to provide efficient gene editing to enable treatment for cancers as well as infectious and genetic disorders [3, 4]. Moreover, genome editing tools offer new opportunities in basic cancer research and diagnosis, including wide advantages such as simple design, rapid operation, low cost, and robust scalability, introducing CRISPR/Cas as a rapidly evolving editing approach that is applicable to almost all genomic targets [5,6,7]. Historically, the term “CRISPR” was proposed by Mojica and Ruud Jansen (2001) [8]; such palindromic repeats were first recognized in Escherichia coli by Ishino et al. (1987) [9]. The function of these sequences remained unclear until 2005. Mojica et al. (2005) first stated that CRISPR serves a significant role in the bacterial immune system [10]. Molecular reports have shown that CRISPR repeats could be detected in around 40% of bacteria and about 90% of archaea [11].

During the last two decades, oncogenes, tumor suppressor genes, metabolism-related genes, and genes involved in resistance to chemo- and radiotherapy have been targeted and edited by using the CRISPR/Cas9 system to constrain tumor growth and progression [12,13,14,15]. Moreover, CRISPR/Cas9-mediated genome editing has wide-ranging potential in cancer therapy. Tumorigenesis is a complicated process including complex interactions between cancer cells and the host immune system [16]. Integration of CRISPR/Cas technique with cancer immunotherapy, such as chimeric antigen receptor (CAR)-T cell-based therapy, and its ability to alleviate carcinogenic viral infections such as human papillomavirus (HPV) has recently emerged as a promising therapeutic approach to a wide range of diseases [17, 18]. Nonetheless, the off-target activity of the CRISPR/Cas9 genome editing tool has been a significant drawback [19]. Therefore, improving its specificity to overcome such off-target effects for safe therapeutic application of CRISPR/Cas9 is of great importance. This study emphasizes recent findings concerning the application of CRISPR/Cas9 (type II CRISPR/Cas9) methods in cancer therapy. We also discuss existing hurdles and contemplate future directions. Furthermore, a glimpse of the ability of the CRISPR/Cas9 system to evolve “off-the-shelf” CAR-T cells with higher anticancer competence is also presented.

CRISPR/Cas9 systems

The growth of artificially designed meganucleases (homing endonucleases) followed by ZFNs and TALENs, and CRISPR/Cas9 has promoted the efficacy of gene editing tools, providing groundbreaking developments in site-specific nuclease (SSN) systems [3]. However, the main drawbacks of cloning and engineering of ZFNs and TALENs have limited their application by the scientific community [20]. In this light, CRISPR technology has renewed SSN systems, resulting in deep editing efficacy and simplicity even for minimal sequences and thus becoming a preferred tool for various genome-targeting goals [21, 22].

Action mechanism

It is now known that bacteria catch snippets of DNA from invading viruses and integrate them into their genome to generate CRISPR arrays, enabling bacteria to become familiar with viruses for their next possible encounter. In response to a subsequent invasion, the bacteria use RNA fragments from such CRISPR arrays to affect the DNA of the viruses [23]. The bacteria then exploit Cas9 or a similar enzyme (e.g., Cas3 and Cas10) to cut the DNA segment, thereby limiting the viability and dangerous functions of the virus. Mechanically, the natural CRISPR/Cas9 system in bacteria consists of two main RNA segments: mature CRISPR RNA (crRNA) and trans-activating crRNA (tracrRNA) [24, 25]. A functional guide RNA (gRNA) is produced by pairing the tracrRNA base with the crRNA. The crRNA sequence can be separated into guide and repeat regions, whereas the tracrRNA sequence includes an anti-repeat region and three stem-loop assemblies. The guide region yields the gRNA:DNA heteroduplex by Watson and Crick base pairing with the DNA target site. The repeat and anti-repeat regions establish the repeat:anti-repeat duplex by Watson and Crick base pairing [24, 26]. Notably, while Cas9 applies the tracrRNA part of the guide as a handle, the crRNA spacer segment directs the complex for identifying viral sequences [27]. Indeed, crRNA and tracrRNAs form Cas9 protein–RNA machinery that cuts the viral sequence with DNA double-strand breaks (DSBs). One of the advantages of this two-component system is that the gRNA can be altered independently from the Cas nuclease, facilitating the modification of CRISPR as a genome editing tool with unrestricted target capability and high efficiency [28, 29]. In contrast to conventional tandem repeats in the genome, CRISPR repeat clusters are separated by nonrepeating DNA sequences termed spacers belonging to dangerous viruses [10, 30]. There is substantial similarity between the spacer sequences and the protospacer-adjacent motif (PAM) sequences targeted by the guide RNA [31]. PAMs are short DNA sequences (typically 2–6 base pairs in length) situated 3–4 nucleotides downstream from the cleavage site. Streptococcus pyogenes Cas9 (SpCas9) nuclease is directed by a sgRNA to a 20-bp sequence of target DNA located next to a three-base-pair PAM (5′-NGG-3′), providing a blunt-ended DNA double-strand break (DSB). DSBs stimulate cellular repair systems, chiefly nonhomologous end-joining (NHEJ, imprecise repair) or homology-directed repair (HDR, precise repair) (Fig. 1). In this regard, CRISPR has become known as a powerful, reprogrammable genome editing tool. CRISPR technology includes an endonuclease such as Cas9 protein concomitant with a single sgRNA, which is functionally comparable to the crRNA-tracrRNA complex in bacteria. The sgRNA plays a paramount role in determining the specificity and cutting activities of the endonuclease [32,33,34].

Fig. 1
figure 1

Action mechanism of CRISPR/Cas9 system, including nonhomologous end-joining (NHEJ), homology-directed repair (HDR), single-guide RNA (sgRNA), and protospacer adjacent motif (PAM)

Classes and types

A series of Cas9 variants have been industrialized to improve the editing fidelity or targeting range of CRISPR/Cas9 (Table 1). Regarding the organization of the effector protein and the presence or lack of signature genes, conservation of the protein sequence, and organization of respective genomic loci, CRISPR systems can mainly be classified into 2 main classes, 6 types, and over 30 subtypes [8]. Class 1 consists of type I and type III CRISPR systems and is typically found in Archaea, while class 2 includes type II, IV, V, and VI CRISPR systems [35, 36]. Class 2 includes only one effector protein, while class 1 comprises multi-subunit Cas protein complexes. Importantly, a unified classification of these systems should be based on various criteria because of the complexity of the genomic architectures and the rapid evolution of CRISPR/Cas systems [36]. Significantly, three specific signature genes distinguish the three central CRISPR systems: Cas3 in type I systems, Cas9 in type II, and Cas10 in type III. Despite the introduction of several CRISPR/Cas systems for gene editing applications, the most broadly used type is the type II CRISPR-Cas9 system from S. pyogenes. In addition, Cpf1 protein derived from AsCpf1 (Acidaminococcus sp.) and LbCpf1 (Lachnospiraceae bacterium) has attracted increasing attention [37, 38]. In general, class 2 systems have more capacity to improve genome editing and genetic screening, as confirmed by several reports using the Cas9 (Csn1), Cas12a (Cpf1), Cas13a (C2c2), and Cas13b (C2c6) systems.

Table 1 Cas9 variants

CRISPR/Cas9 applications in viral infections and genetic disorders

Viral infection

The CRISPR/Cas9 tool can be applied not only to modify particular nucleotide sequences in the human genome but also to target the double-stranded DNA (dsDNA) of viruses [39]. Interestingly, the CRISPR/Cas9 machinery can be equipped with multiple sgRNAs, which facilitates action on various genomic loci in a single cell by Cas9 endonucleases [40, 41]. Cas9 variants also enable targeted gene mutation, transcriptional activation and suppression, epigenetic alteration, imaging of DNA loci, and single-base mutations [28, 42]. By using the CRISPR/Cas9 system, clearance of viruses from infected cells becomes hypothetically practical for any DNA- or RNA-mediated virus during their pathological process. Therefore, the CRISPR/Cas9 technique has become a game-changing tool for modifying several developmental phases of the viral life cycle and holds the capacity to enable efficient genetic therapy versus human viruses (Table 2) [43, 44]. In recent years, CRISPR/Cas9-mediated antiviral protocols to manipulate infectious human viruses have been applied efficiently. In this regard, the CRISPR/Cas9 system has shown remarkable efficacy against human immunodeficiency virus (HIV), hepatitis B virus (HBV), and HPV [45, 46].

Table 2 CRISPR/Cas9 applications in treatment of infectious disease

The most promising editing targets of CRISPR/Cas9 therapy to combat HIV viruses are the C–C chemokine receptor 5 (CCR5) gene, C–C–C chemokine receptor 4 (CXCR4) gene, proviral DNA-encoding viral proteins, and the HIV 5′ and 3′ long terminal repeat (LTR) [47,48,49]. For instance, Ebina et al. (2013) showed the extraordinary capacity of the CRISPR/Cas9 machinery to affect the HIV-1 genome and avert its expression [50]. They found that LTR-targeting CRISPR/Cas9 reagents suppressed LTR-driven expression in HIV-1-infected T cells and cleaved and mutated LTR target sites, leading to perturbation of latent HIV-1 provirus [50]. Also, Cas9-induced ablation of CXCR4 in T cells resulted in their robust resistance to HIV without significant off-target effects and disrupting cell biological processes such as proliferation [51]. In addition to non-carcinogenic viruses (e.g., HIV), the CRISPR/Cas9 system offers the opportunity to modify the pathogenic process of carcinogenic viruses such as HPV and HBV. Viruses are causal agents of about 10–15% of all cancers worldwide in addition to their prominent role in infectious diseases. Among viruses, several DNA viruses, including Kaposi’s sarcoma herpesvirus (KSHV), Epstein–Barr virus (EBV), HPV, HBV, and simian virus 40 (SV40), along with two RNA viruses, viz. human T-lymphotropic virus-1 (HTLV-1) and hepatitis C virus (HCV), are the most well-defined carcinogenic viruses [52, 53]. Zhen and colleagues (2015) suggested that CRISPR/Cas9-mediated ablation of the surface antigen (HBsAg)-encoding region of HBV prohibited HBV replication in liver-derived cell lines, HepG2, and BALB/c nude mice, as evidenced by reduced levels of HBsAg secretion in cell culture and mouse serum [54]. Likewise, the CRISPR/Cas9 system targeted HBV covalently closed circular DNA (cccDNA) and inhibited HBV replication in HBV-infected Huh7 and HepG2 cells. By means of the CRISPR/Cas9 system, clearance of viruses from infected cells becomes hypothetically practical for any DNA- or RNA-mediated virus during their pathological process [55]. Thus, the CRISPR/Cas9 system may serve as a unique avenue for HBV therapy. In addition to the use of the CRISPR/Cas9 system alone, combination therapy of CRISPR/Cas9 with other modalities, such as the NU7026 P inhibitor, could efficiently eliminate the HBV genome from infected cells [56]. NU7026 P is a well-known suppressor of NHEJ and constrains CRISPR/Cas9-mediated degradation of cccDNA and results in large on-target deletions [56]. Thus, negative regulation of its activation may potentiate the efficacy of CRISPR/Cas9-mediated degradation of cccDNA, culminating in HBV genome eradication. Given the central role of early genes E6 and E7 in continuing the malignant phenotype of cervical cancer cells following HPV infection, the CRISPR/Cas system has recently been applied to target HPV16/18-E6 and-E7 DNA in HPV-infected cells [57, 58]. In this regard, HPV16-E7 ablation using the CRISPR/Cas system induced apoptosis and disrupted proliferation of cervical cancer SiHa and Caski cell lines in vitro with no effect on HPV-negative cells [59]. The E7 DNA deficiency resulted in upregulation of tumor suppressor protein retinoblastoma (pRb), suggesting E7 as a potential target for gene editing approaches to treat cervical cancer [59]. Furthermore, it has been shown that addition of the CRISPR/Cas9 system to immune checkpoint inhibitors (ICIs), as Food and Drug Administration (FDA)-approved anticancer drugs, may augment their antitumor effects [60,61,62]. For example, combination therapy using CRISPR/Cas9-mediated disruption of HPV16 E6/E7 gene and PD1 inhibitor resulted in an improved overall survival (OS) rate accompanied by impaired tumor development in SiHa tumor cell-bearing SCID mice [63, 64]. Also, cotreatment inspired the population of antigen-presenting cells (APCs), CD8+ and CD4+ T lymphocyte cells in tumor tissue, thereby eliciting robust antitumor responses in treated mice against tumor tissue [63].

A growing body of evidence indicates that the CRISPR/Cas9-mediated genetic targeting tool could be an alternative means to treat virus-related diseases in the future. Nonetheless, viruses can evade CISPR/Cas9-mediated inhibition by attaining several mutations at the target region, which disfavors gRNA interaction with the corresponding sequence, without deterring viral replication [65, 66]. Circumventing this drawback is thus urgently required before their wide application in the clinic.

Genetic disorders

Gene targeting systems have provided a quick and operational means to target and modify the genome at specific sites. Many genes contribute to the pathogenesis of genetic disorders [67, 68]. Given that one particular genetic mutation causes such genetic disorders, the CRISPR/Cas9 machinery can be used to treat such disorders by targeting and modifying a single gene [69, 70]. Such targeting of genes can be accomplished both ex vivo and in vivo [71]. The target cells with mutated genes are isolated then manipulated by programmable nucleases to correct the mutated gene, and ultimately injected into the original host ex vivo [72, 73]. Engineered nucleases accompanied with the correct sequence of the target gene can be injected directly into the patient for systemic or targeted tissue (such as the eye, brain, or muscle) in vivo [74, 75]. During the recent decade, CRISPR/Cas9 has exhibited promising preliminary capability to treat β-thalassemia [76,77,78], tyrosinemia [79], Duchenne muscular dystrophy (DMD) [80, 81], hemophilia [82, 83], cystic fibrosis [84], central nervous system (CNS)-associated diseases [85, 86], Tay–Sachs diseases (TSD) [87], and fragile X syndrome disorders (FXS) [88, 89]. Indeed, this technology has enabled the correction of the multiple mutated genes associated with responding genetic disorders, including the DMD gene in DMD, CFTR gene in CF, factor IX gene in hemophilia B, hemoglobin beta-chain gene in β-thalassemia, presenilin 1 and 2 (PSEN1 and PSEN2) and apolipoprotein E4 (apoE4) genes in AD, HTT gene in HD, leucine-rich repeat kinase 2 (LRRKK2) gene in PD, fumarylacetoacetate hydrolase (FAH) in tyrosinemia, Hex gene in TSD, fragile X mental retardation 1 (FMR1) gene in FXS, etc. [90, 91].

A complete review of such CRISPR/Cas9 applications lies beyond the scope of this article, so the reader is referred to excellent articles in this field [92,93,94].

CRISPR/Cas9 in cancers

CRISPR/Cas9 tools have great capacity for the diagnosis and treatment of cancer, including (1) the use of CRISPR/Cas9-based diagnostic systems SHERLOCK and DETECTR for cancer diagnostics, (2) providing TCR knockout (KO) CAR-T cells (universal CAR-T cells), (3) KO of inhibitory receptors such as PD-1 and LAG-3 to promote the capability of cancer immunotherapy, (4) elimination of oncogenic virus-like HPV, (5) and establishment of in vivo tumor models by eliciting mutations in several genes [7, 45, 95,96,97] (Fig. 2).

Fig. 2
figure 2

CRISPR/Cas9 applications in cancer research and therapy. Knockout (KO), T-cell receptor (TCR), chimeric antigen receptor (CAR)-T cell, β2-microglobulin (B2M), programmed cell death protein 1 (PD1 or PDCD1), lymphocyte activating gene 3 (LAG-3), transforming growth factor-beta receptor (TGF-βR), diacylglycerol (DAG), Epstein–Barr virus (EBV), human papillomavirus (HPV), hepatitis B virus (HBV), hepatitis C virus (HCV), cancer stem cell (CSC)

In this section, we focus on the therapeutic potential of the CRISPR/cas9 system in cancer treatment (Table 3).

Table 3 CRISPR/Cas9 applications in cancer treatment

Liver cancer

Recently, targeting of various genes in liver cancer cells using the CRISPR/Cas9 system has demonstrated a potential ability to impair their proliferation and metastasis. In 2019, Zhang and colleagues designed a specific sgRNA to target nuclear receptor-binding SET domain-containing protein 1 (NSD1) in HCC cell lines [98]. The NSD1 histone lysine methyltransferase targets the Wnt/β-catenin signaling pathway associated with HCC tumorigenesis. They found that CRISPR/Css9-mediated NSD1 KO HCC cells displayed reduced proliferation, migration, and invasion in vitro and in vivo [98]. NSD1 ablation brought about improved methylation of H3K27me3 and reduced methylation of H3K36me2, leading to downregulation of Wnt10b expression. Therefore, the CRISPR/Cas9 tool may hinder HCC oncological events by negatively regulating the Wnt/β-catenin signaling axis in nude mice and in vitro [98]. In HCC, targeting the Wnt/β-catenin signaling axis using CRISPR/Cas9 machinery could exert a positive antitumor effect in HEK 293T cell line, as evidenced by their perturbed proliferation [99]. Likewise, CRISPR/Cas9-mediated ablation of acid-sensing ion channels 1a (ASIC1a), which triggers migration and invasion in liver cancer, could deter cell proliferation and tumorigenicity [100]. The empirical consequences are caused mainly by stimulation of β-catenin degradation and coactive lymphoid enhancer factor/T cell factor (LEF/TCF) inactivation in HCC cell lines and also xenograft mice following ASIC1a ablation [100]. Mechanistically, the β-catenin inspires downstream signaling transduction by LEF-TCF, which eventually induces c-MYC expression [101]. In malignant cells, the β-catenin/LEF/TCF axis is often prompted and triggers cell proliferation [102]. Besides, inhibition of this pathway may offer great potential to moderate HCC proliferation, migration, and invasiveness [103]. Also, dysregulation of insulin-like growth factor 2 (IGF2) mRNA-binding protein 1 (IGF2BP1) has been suggested to be involved in HCC progression [104]. IGF2BP1 is required to stabilize and translate various oncogenes, such as glioma-associated oncogene homolog 1 (Gli1) and Myc 91, thus its expression is associated with worse prognosis in HCC patients [104, 105]. LIN28B-AS1 directly binds to IGF2BP1 like long non-coding RNAs. Zhang et al. (2020) designed specific sgRNA targeting and modified LIN28B-AS1 expression to evaluate LIN28B-AS1 ablation on HCC proliferation pathological events [106]. They found that IGF2, Gli1, and Myc expression was substantially downregulated in LIN28B-AS1-deficient HCC cell lines in vitro by reducing IGF2BP1 mRNA levels, suppressing HCC cell proliferation and invasion [106]. In nude mice, LIN28B-AS1 KO HepG2 xenograft tumors had a slightly increasing trend compared with normal LIN28B-AS1-positive HepG2 xenograft tumors [106]. In addition to gene editing tools, negative regulation of IGF2BP1 synthesis in HCC cells using specific siRNA dissuades tumor proliferation and invasion [107]. These findings make IGF2BP1 a potent target for HCC therapy with the aim of delivering novel therapeutic plans with improved efficacy.

Angiogenesis plays a fundamental role in tumor progression. Meanwhile, the angiopoietin-2 (ANGPT2)/Tie2 pathway induces angiogenesis in HCC tumors by directly targeting the proliferation of endothelial cells [108, 109]. Accordingly, the ANGPT2/Tie2 axis has been suggested as a reasonable target for antiangiogenic therapy. Targeting ANGPT2 is presently undergoing phase II clinical trials, with preliminary results suggesting encouraging antitumor activity and safety [110]. In 2020, Xie and colleagues showed that CRISPR/Cas9-mediated ablation of ANGPT2 in Hep3B and MHCC97H cell lines diminished the potential of their derivative exosome to promote proliferation of ECs [109]. It thus appears that this pathway could be a putative therapeutic target for antiangiogenic treatments. Also, the role of miR-3188 in HCC pathogenesis has recently been manifested, where its overexpression improves cell viability and proliferation but suppresses apoptosis of HCC cells [111, 112]. Meanwhile, Zhou et al. (2017) showed that miR-3188 ablation could constrain cell growth and colony formation, induce cell cycle arrest (G0/G1 phase), and instigate apoptosis in HepG2 cells [113]. miR-3188 inactivation could also diminish migration and invasion due to downregulation of Notch1 activation in HCC cells [113].

Importantly, gene editing tools can defeat HCC resistance to conventional treatments such as sorafenib therapy. Because of the central role of phosphoglycerate dehydrogenase (PHGDH), which serves a critical role in serine synthesis and triggering HCC resistance to sorafenib, Wei et al. (2019) highlighted its potency to compromise Sorafenib resistance [114]. They showed that downregulation of nicotinamide adenine dinucleotide phosphate (NADPH) enforced PHGDH KO HCC cells to increase reactive oxygen species (ROS) levels. Manipulated cells also showed higher apoptosis rates upon sorafenib treatment than nonmanipulated cells [114]. It was proposed that PHGDH ablation results in negative regulation of the synthesis of antioxidant mediators (e.g., NADPH) and then makes PHGDH KO HCC cells susceptible to sorafenib [114].

Colorectal cancer (CRC)

The latest investigation has revealed that the CRISPR/Cas technique could target long non-coding RNAs (lncRNAs), thus enabling CRC treatment. Researchers have sought different strategies to suppress their activity to achieve better therapeutic outcomes. For instance, CRISPR/Cas9-mediated ablation of lncRNA CCAT1 gene in other CRC cell lines, SW-480 and 14 HCT-116, could impair their anchorage-independent growth [115]. CCAT1 expression has an intimate association with the CRC stage and stimulates cell growth and mobility by targeting miR-181a-5p [116]. Therefore, it may be possible to target CRC therapy due to its undesired biological activities. Likewise, in mouse and human tumor-derived organoids, simultaneous targeting of adenomatous polyposis coli (APC) and KRAS, which mainly contribute to the disease progress in the early stage of CRC, brought about robust antitumor effects [117]. In addition, secretory mucin (MUC) 5AC has recently been suggested as a putative target for targeted therapy [118, 119]. MUC5AC is a large gel-forming glycoprotein expressed aberrantly during CRC stages [120]. A recent study in subcutaneous and colon orthotopic mouse models demonstrated that MUC5AC-deficient CRC cells possess less tumorigenic capacity [121]. Also, MUC5AC-deficient tumor-cell-bearing mice exhibit reduced appearance of metastatic lesions [121]. Since MUC5AC induces chemical resistance through CR44/β-catenin/p53/p21 signaling in CRC 107, combination therapy with gene editing tools and chemotherapeutic agents can break CRC resistance to conventional chemotherapies [121]. In addition to MUC5AC, it has been proposed that ablation of partitioning defective 3-like protein (Par3L) 108, a recently described cell polarity protein, and nuclear factor-erythroid factor 2-related factor 2 (Nrf2) [122], a critical transcription factor, may attenuate CRC cell resistance to chemotherapies and irradiation. Mechanistically, Par3L plays a crucial role in CRC survival via negative regulation of the liver kinase B1 (LKB1) Lkb/AMP-activated protein kinase (AMPK) signaling pathway [123]. Besides, NRF2 potentiates amino acid and protein synthesis in CRC cells [122], so targeting its expression could result in encouraging outcomes in CRC. Another study applied the CRISPR/Cas9 technique to target dachshund homolog 1 (DACH1), a target expressed explicitly in discrete crypt base cells [124]. DACH1 protein promotes tumorigenesis, invasion, and metastasis by deregulating the bone morphogenetic protein (BMP) signaling pathway [125, 126]. Importantly, its levels are usually found to be boosted in all stages of CRC [127]. Nonetheless, KO of DACH1 expression using CRISPR technique and shRNA could deter CRC cell growth, attenuate organoid formation efficiency, and organoid tumor size [127]. These results shed light on the role of DACH1 and introduce a possible prognostic marker and therapeutic goal for CRC patients [127]. Furthermore, lysine-specific demethylase 1 (LSD1), a well-known chromatin-modifying enzyme, is overexpressed in CRC and associated with proliferation and migration mainly by transduction of the phosphoinositide 3-kinase (PI3K)/Akt axis [128]. In this regard, Miller et al. applied specific sgRNA to block its expression in CRC cell lines and showed that CRISPR/Cas9-mediated LSD1 ablation corresponded to inhibition of AKT-induced epithelial–mesenchymal transition (EMT) and migration [129]. Other studies have outlined that inactivation of LSD1 by gene editing techniques could inhibit the proliferation and migration of leukemia [130, 131], Merkel cell carcinoma (MCC) [132], and HCC cells [133].

Breast cancer

It has been strongly evidenced that altered expression of miRNAs is involved in breast cancer progression [134, 135]. In this regard, miR-23b and miR-27b promote tumor progress in various human tumors and may provoke the angiogenesis process in this setting. Recent studies in MCF7 breast cancer cells demonstrated that KO of miR-23b and miR-27b gene expression using CRISPR systems alleviated tumor growth in xenograft nude mice by upregulation of ST14 (suppression of tumorigenicity 14) [136]. ST14 typically decreases breast cancer cell proliferation and invasion [137, 138], so antitumor effects upon inactivation of miR-23b and miR-27b may depend on promotion of ST14 activity. Increasing evidence also shows that dysregulated expression of fatty acid synthase (FASN), complicating the endogenous synthesis of fatty acids and the adjustment of ERα signaling, may contribute to breast cancer onset and progress [139]. In 2020, Gonzalez-Salinas et al. showed that CRISPR/Cas9-mediated genetic depletion of FASN inhibits aggressive features in breast cancer MCF-7 cells, as verified by impaired cell proliferation, viability, and migration [140]. Importantly, transcriptomic studies have revealed that FASN deficiency has a more evident negative effect on proliferation-associated genes than lipid metabolism [140]. These results were also confirmed by analysis of the impact of FASN KO on oncogenic activities in leukemia cells [141].

Furthermore, targeting platelet glycoprotein VI (GPVI), which acts as a metastasis inducer by interaction with cancer cell-derived galectin-3, resulted in marked antitumor activities in vitro and in vivo [142]. GPVI causes the maintenance of tumor vessel integrity and mediates interactions between platelets and cancer cells. Platelets protect cancer cells from attack by natural killer cells (NKCs) [143], so perturbing platelet–cancer cell interaction may disrupt tumor cell progress. In this regard, Mammadova-Bach and coworkers (2020) reported that KO of platelet GPVI in mice led to a drop in breast cancer cell metastasis [144]. Also, GPVI inhibitors were found that could provide an obstacle to ovarian [142] and prostate [145] cancer metastasis. GPVI may thus be a potential target for antimetastatic treatments. Also, impaired breast cancer cell proliferation and metastasis were observed following CRISPR/Cas9-mediated inactivation of fucosyltransferase 8 (FUT8), a critical positive regulator of cell growth and tumor metastasis core fucosylation of target biomolecules [146]. FUT8 ablation alleviates TGF-β signaling and EMT in breast cancer by inhibiting TGF-β core fucosylation, disturbing breast cancer lung metastasis in mice xenografts [146].

Since increased expression of doublecortin-like kinase 1 (DCLK1) has been reported in patients with breast cancer associated with poor prognosis, targeting DCLK1 has been proposed as a possible candidate in the field of antitumor study [147]. Liu and coworkers (2019) found that DCLK1 KO in breast cancer cell line BT474 using CRISPR technology suppressed its metastatic features [148]. These beneficial effects were likely related to upregulation of tight junctions (TJ)-associated protein Zonula occludens (ZO-1) along with downregulation of zinc-finger E-box binding homeobox 1 (ZEB1), a master regulator of EMT [148]. Indeed, upregulation of TJ-associated protein expression and conversely suppressing ZEB1 activation, in turn, leads to reduced cell motility and invasiveness [148]. Gene editing tools can offer a practical possibility for overcoming cancer cell resistance to conventional therapies. Due to the proven role of the osteopontin (OPN) gene in inducing resistance to radiotherapy (RT) [149, 150], the impacts of its ablation in conjunction with RT have been highlighted [151]. Accordingly, Behbahani et al. (2021) indicated that the viability of the OPN-deficient breast cancer MDA-MB-231 cell line was severely reduced upon RT compared with the nonmanipulated MDA-MB-231 cell line [151]. It can thus be supposed that inactivation of the OPN gene might become an effective therapeutic plan to circumvent tumor cell resistance to conventional therapies, such as RT [151].

Cervical cancer

Targeting oncoproteins E6 and E7 in HPV16 and HPV18 utilizing gene editing tools could inactivate such oncogenes and thus prompt cell cycle arrest and apoptosis [58, 152]. For example, Ling et al. (2020) showed that double-targeting of E6 and E7 improved p53 and p21 protein levels in cervical cancer lines (HeLa and SiHa) and tumor cell-bearing mice [58]. Given that HPV E6 stimulates inactivation of p53 in tumor cells, reactivation of its expression and transduction of p53 signaling upon E6 ablation has been recommended as a putative scheme for cervical cancer therapy [153]. In addition to the CRISPR system, ZFNs- [154] and TALEN-based [155] targeting of HPV16/18 E7 could efficiently block expression of E7 oncogenes and lead to apoptosis induction in HPV16 HPV18-infected cervical cancer cells.

The latest research has shown that targeting aldo–keto reductase family one member B1 (AKR1B1), which is highly expressed in several tumors and correlates with tumor growth, could benefit cervical cancer [156]. AKR1B1 contributes to prostaglandin F2α (PGF2α) synthesis and protein kinase C (PKC) transduction, which in turn triggers upregulation of NF-kB, inflammation, and inflammation proliferation [157]. Improved AKR1B1 levels and potentiated activity are usually detected in cervical cancer, which hypothetically correlates with higher prostaglandin E2 (PGE2), a well-known inducer of cervical carcinogenesis [156, 157]. The establishment of human endometrial KO cell lines using CRISPR/Cas9 technology confirms the PGs synthase function of AKR1B1 [158]. In vitro studies have shown that AKR1B1-deficient cervical cancer cell lines exhibit lower proliferation, migration, and invasion than nonmanipulated cells [159]. Concerning recent reports, AKR1B1 suppression could constrain PGE2 activity and thus disturb cervical carcinogenesis by preventing angiogenesis and cancer cell proliferation as well as inducing apoptosis [160]. Also, CD109, as a result of its role in transforming growth factor-β1 (TGF-β1) signaling and signal transducer and activator of transcription 3 (STAT3) activation, could be an innovative target for cervical cancer therapy [161,162,163]. CD109 is drastically expressed in cervical cancer and upregulates epidermal growth factor receptor (EGFR)-mediated STAT3 phosphorylation, enabling cervical cancer cell migration and proliferation, and supporting cancer cell phenotype [161]. However, Mo et al. (2020) demonstrated that targeting CD109 by siRNA or CRISPR/Cas9 could inhibit cervical cancers’ tumorigenic and aggressive properties by inactivating the CD109/EGFR/STAT3 axis in vitro and in vivo [161].

Furthermore, KO of growth differentiation factor-8 (GDF-8), or myostatin, a protein that is highly overexpressed in human tumors, by using the CRISPR/Cas9 technique could induce apoptosis intrinsic pathways in HeLa cells and prohibit their proliferation [164]. The observed effects are probably caused by increased ROS intracellular levels and promotion of elevated fatty acid oxidation, which leads to induction of mitochondrial membrane depolarization, secretion of cytochrome c (Cyt-c), and finally induction of the caspase cascade [164]. Similarly, targeting GDF-8 expression in Lewis lung carcinoma (LLC) cells impaired their proliferation and growth in vitro and in vivo [165]. Also, KO of GDF-8 promotes skeletal muscle mass in tumor-bearing rodents through upregulation of the Akt/mTOR pathway, easing the production of skeletal muscle proteins [165].

Lung cancer

Recent studies have highlighted the role of YES1 in lung cancer development, identifying YES1 as a potential target involved in lung cancer carcinogenesis [166]. YES1 adjusts cell growth, survival, apoptosis, cell–cell adhesion, and cytoskeleton remodeling. Its levels have been found to be enhanced in patients with lung cancer, making it a potential therapeutic target in lung cancer [167]. The vital role of YES1 in lung carcinogenesis was revealed by its obstruction using the CRISPR/Cas9 system, leading to disrupted growth and metastasis of NSCLC by downregulation of mTOR signaling, a positive regulator of carcinogenesis [168]. Also, genetic depletion of YES1 made dasatinib-resistant NSCLC cell lines susceptible to dasatinib-induced antitumor effects in vitro [168]. Its congenital absence also led to promising impacts in other malignancies, such as breast [169] and ovarian cancers [170]. Moreover, Zhang et al. (2020) evaluated the possible effect of the genetic depletion of Fyn-related kinase (FRK) by CRISPR/Cas9 in lung carcinoma H1299 cells to elucidate its role in NSCLC pathogenesis [171]. FRK potentiates the stemness phenotype of NSCLC and triggers the EMT process by eliciting metabolic reprogramming [172, 173].

Interestingly, FRK depletion impaired the stemness phenotype of H1299 by downregulation of CD44 and CD133 expression and concurrently stimulated metabolism reprogramming by blocking the Warburg effect and varying the energy type in H1299 cells [171]. Also, FRK-deficient H1299 cells demonstrated attenuated proliferation, invasion, colony formation, and EMT process in vitro. These findings indicate that FRK could be a putative target for lung carcinoma therapy [171].

In 2020, Grunblatt and colleagues showed that CRISPR/Cas9-mediated KO of oncogene N-MYC may yield small cell lung cancer (SCLC) [174]. Amplification of N-MYC is a well-recognized poor prognostic marker for human tumors and is associated with aggressive tumor features and resistance to conventional therapies [175]. The results of a study conducted in chemosensitive patient-derived xenograft (PDX) models of SCLC revealed that inactivation of N-MYC restores cancer cell chemosensitivity through downregulation of ubiquitin-specific protease 7 (USP7) expression [174]. USP7 favors DNA damage response and stimulates cancer progress by negative regulation of p53, and is associated with poor survival rate in cancer patients [176, 177]. Hence, inactivating its expression using inhibition of N-MYC expression or its direct ablation has been an imperative strategy in cancer therapy [176, 178].

Pancreatic cancer

KRAS mutation has been confirmed as the primary contributor to pancreatic cancer carcinogenesis, being mutated in ~ 95% of pancreatic neoplasias [179]. In 2019, Lentsch et al. found that efficient KO of c.35G > A (p.G12D) Kras mutation in human pancreatic cancer cell lines SUIT-2 and Panc-1 and mouse cell lines TB32047 is possible [180]. Studies in pancreatic ductal adenocarcinoma (PDA) rodent models indicated that KRAS favors immune escape in pancreatic cancer cell-bearing mice by activating the BRAF and MYC axis [181]. However, KRAS genetic depletion using the CRISPR system provokes an antitumor response against PDA cells. Of course, KRAS ablation attenuates, but does not eliminate, the tumorigenic potential of PDAC cells, suggesting that the multifaceted axis complicates the progress of PDA [181]. Given that the hypoxic tumor microenvironment (TME) supports the growth and metastasis of pancreatic cancer cells [182], inactivation of hypoxia-inducible factor-1α (HIF-1α) with CRISPR/Cas9 is suggested as another rational therapeutic approach [183]. For the first time, Li et al. (2019) developed a tumor-targeted lipid-based CRISPR/Cas9 delivery system to inhibit HIF-1α expression in vitro and in vivo [184]. They showed that ablation of HIF-1α resulted in lower expression of its downstream targets such as vascular endothelial growth factor (VEGF) and matrix metalloproteinase-9 (MMP-9), ensuring reduced metastasis and ameliorating the paclitaxel-driven cytotoxicity on human pancreatic cancer cell line BxPC-3 in vitro and in vivo [184]. It appears that combining CRISPR technology with conventional therapies could be a more efficient antitumor strategy. Likewise, Wei and colleagues (2020) revealed that ablation of protein arginine methyltransferase 5 (PRMT5), a central transcriptional regulator, by using the CRISPR/Cas9 technique enhances the susceptibility of PDAC cells to gemcitabine by inducing cell cycle arrest [185]. Other studies have signified that targeting autophagy may affect aggressive features of pancreatic cancer [186]. Meanwhile, Hwang and coworkers (2019) evaluated the role of EI24 (etoposide-induced gene 2.4 kb; PIG8, p53-induced gene 8) as a component of autophagy in pancreatic cancer cell growth [187]. They found that knockdown (KD) or KO of EI24 utilizing siRNA or CRISPR/Cas9, respectively, impaired pancreatic cancer autophagy and then suppressed cell proliferation [187]. These results indicate that EI24 acts as a tumor inducer in pancreatic cancer cells; however, there are some conflicting reports. For example, Zang et al. (2018) described that EI24 inhibits cell proliferation and stimulates cell cycle arrest in PDAC cells by triggering autophagic lysosomal degradation of c-Myc proto-oncogene [188]. Therefore, further analysis of the data and execution of more comprehensive studies are required to clarify the detailed role of EI24 in pancreatic cancer carcinogenesis.

Prostate cancer

Many studies have revealed that activating protein-1 (AP-1), a transcription factor, is related to cancer onset and progress [189]. The proto-oncogenes JUN and FOS are pivotal in prostate cancer progression and invasion [190]. In prostate cancer cells, Ouyang et al. (2008) showed that forced expression of c-Fos and c-Jun stimulates tumorigenicity and provokes transduction of ERK/MAPK signaling [191]. Besides, Riedel and coworkers (2021) evidenced that CRISPR/Cas9-mediated inactivation of Jun results in impaired prostate cancer cell proliferation and invasiveness in vitro and in vivo [192]. Also, ablation of FOS potentiates Jun expression, and CRISPR/Cas9-mediated KO of Jun constrains prostate cancer cell proliferation [193]. Hence, targeting AP-1 transcription factors in prostate cancer by genome edition could be a therapeutic approach. In addition to AP-1, G protein-coupled receptor family C group 6 member A (GPRC6A) as a functional osteocalcin and testosterone sensing receptor contributes to prostate cancer growth [194]. In this regard, its upregulation enables prostate cancer cells to grow in response to dietary and bone-derived ligands [194]. Although it induces the EMT process of prostate cancer, KD of GPRC6A attenuates such cell invasion [194]. Importantly, GPRC6A-deficient prostate cancer cell line PC-3 created by CRISPR/Cas9 technology demonstrates drastically lower growth and aggression than nonmanipulated cells in vitro and in vivo [195]. Also, manipulated cells showed reduced ligand-dependent responses in vitro due to downregulation of extracellular-signal-regulated kinase (ERK) activity [195]. In another study, Chakraborty et al. (2021) designed a specific sgRNA to target expression of BRCA2, a key component of DNA damage repair (DDR). Its mutations have a tight association with prostate cancer oncological events [196]. Variations in DDR pathway genes such as BRCA1/2 and ATM occur in 20–25% of men with metastatic castration-resistant prostate cancer (mCRPC) and complicate cancer cell resistance to therapeutic modalities [196]. They found that genetic depletion of BRCA2 established by the CRISPR system caused an antiproliferative effect on prostate cancer cells and enhanced their susceptibility to poly (ADP-ribose) polymerase (PARP) inhibitors, FDA-approved drugs for mCRPC treatment [196]. Finally, double KO of Akt1 and Akt2 genes potently decreased prostate cancer cell metastasis in vitro and in vivo [197]. Aberrant expression of Akt1 and Akt2 with poor prognosis is shown in various cancers, such as colon [198], gastric [199], breast [200, 201], NSCLC [202], ovarian [203], HCC [204], and pancreatic cancers [205]. Indeed, Akt promotes cell survival, metastasis, and angiogenesis by downregulation of proapoptotic signals, such as Bad and Forkhead box O (FOXO) transcription factors, and transducing VEGF signaling axis [206,207,208]. Interestingly, Su et al. (2021) exhibited that Akt1- and Akt2-deficient prostate cancer CWR22rv1 cells exhibited an enormous invasive reduction in vitro and in vivo [197]. Thereby, inactivation of its expression and activity could offer promising outcomes in vivo.

CRISPR/Cas9 application has also attracted increasing attention for treating other human cancers, such as gastric cancer and glioma [209, 210]. Zhang et al. (2019) recently showed that CRISPR/Cas9-mediated ablation of the prostate-derived Ets factor (PDEF) gene resulted in suppression of the migration and motility of human gastric cancer AGS cells [211]. PDEF as a member of the Ets family of transcription factors serves a key role in stimulating tumorigenesis in gastric cancer, and elevated levels of PDEF correlate with poor prognosis [211]. Thus, targeting its expression could be a putative therapeutic strategy to hinder gastric cancer cell proliferation and metastasis [211]. Likewise, targeting sodium/glucose cotransporters 1 (SGLT1) protein, primarily expressed in various human tumors, is an effective plan to moderate gastric cancer pathogenesis [212, 213]. Its expression is positively related to histological differentiation and worse overall survival in gastric cancer patients [214]. Accordingly, CRISPR/Cas9-mediated ablation of SGLT1 averts proliferation of gastric cancer cells, induces their apoptosis, and could thus modify the metabolism of gastric cancer cells [214]. These results make it a rational target to control the development of gastric cancer cells by influencing their key oncogenic activities. In addition, Haghighi and coworkers (2021) demonstrated that targeting specific genes using genome editing tools could bring about cell cycle arrest in gastric cancer cells [215].

Meanwhile, they found that CRISPR/Cas9-mediated knockout of nuclear paraspeckle assembly transcript 1 (NEAT1) in AGS cells eventually caused S phase cell cycle arrest in vitro [215]. NEAT1, as a lncRNAs, contributes to adjusting cell cycle progression, apoptosis, cell growth, proliferation, and migration in various cells [216, 217]. Also, ablation of NEAT1 triggered apoptosis of AGS cells, in part by upregulation of FAS level, thereby eliciting caspase cascade activation [215]. Besides, other reports have indicated that knockout of the EGFR mutation vIII (EGFRvIII) may target glioma cells’ pathogenesis [218, 219]. It seems that EGFRvIII ablation abrogates NF-κB activation in glioma cells and may thereby improve the overall survival rate in glioma patients [220]. Given the positive association between the expression of tumor vascular laminin-411 (α4β1γ1) with potentiated tumor growth and with the expression of cancer stem cell (CSC) markers, other studies have focused on targeting its expression to assess its role in glioma models [221, 222]. Elevated levels of laminin-411 also have a tight interrelation with increased recurrence rate and shorter survival of glioma patients [223]. Interestingly, KO of the laminin-411 α4 and β1 chains with CRISPR/Cas9 could reduce tumor growth in glioma cell-bearing mice and considerably improve their survival because of downregulation of the Notch pathway [224]. Concerning the assumed hypothesis indicating that Notch signaling can stimulate glioma aggressiveness, targeting up- or downstream of Notch could be a rational approach to alleviate disease progression in vivo [224].

CRISPR/Cas9 in CAR-T cell therapies

Chimeric antigen receptors (CARs) have been applied to genetically engineer T effector cells to potentiate adoptive cellular therapy (ACT) and tumoricidal activities [225]. CARs as recombinant synthetic surface receptors can recognize a specific target antigen on the surface of cancer cells, and subsequently bring about the induction of redirected effector cells activation. The basic CAR construct is made up of a single-chain variable fragment (scFv; ectodomain) that serves as an extracellular antigen-recognition domain [226, 227]. Meanwhile, CAR-T cell therapy has resulted in excellent outcomes in the treatment of a variety of hematological malignancies including acute lymphoblastic leukemia (ALL), chronic lymphocytic leukemia (CLL), lymphoma, and multiple myeloma (MM) [228]. Additionally, CAR-T cell research and development has shown great promise in solid tumors including melanoma, NSCLC, breast cancer, and sarcoma [229, 230].

Despite this groundbreaking success, obstacles to CAR-T cell therapy include three main challenges: (1) the need for case-by-case autologous CAR-T cell generation, (2) cancer cell resistance to CAR-T cell therapy, and (3) occurrence of unwanted toxicities and, more importantly, cytokine release syndrome (CRS) [231]. The need to create autologous CAR-T cells on a case-by-case basis prevents its large-scale clinical application due to the expensive and lengthy manufacturing process [232,233,234]. Induced CAR-T cells could express immune checkpoint molecules such as PD1 and lymphocyte activation gene 3 (LAG3) or CD223, thus deterring CAR-T anticancer function upon interaction with corresponding ligands expressed by cancer cells [235, 236]. Activating a significant number of CAR-T cells concurrently and secretion of higher levels of GM-CSF, IL-6, and IL-1 may bring about CRS [237]. Generating off-the-shelf, allogeneic CAR-T cells with robust resistance to immunosuppressive TME accompanied by lower toxicity is urgently required (Fig. 3) (Table 4).

Fig. 3
figure 3

CRISPR/Cas9 application for manufacture of next-generation CAR-T cells. Knockout (KO), T cell receptor (TCR), chimeric antigen receptor (CAR), human leukocyte antigen (HLA), granulocyte–macrophage colony-stimulating factor (GM-CSF), programmed cell death protein 1 (PD1 or PDCD1), lymphocyte activating gene 3 (LAG-3), transforming growth factor-beta receptor (TGF-βR), diacylglycerol (DAG)

Table 4 CRISPR/Cas9 applications in CAR-T cell therapy

Off-the-shelf or universal CAR-T cells

Recent reports have shown that genetic depletion of T cell receptor (TCR) alpha constant (TRAC or TCR) and β-2 microglobulin (B2M), a component of MHC class I molecules (MHC-1 or HLA-1), by CRISPR/Cas9 may efficiently enable generation of universal CAR-T cells [238]. KO of β2M or TARC impairs allogeneic cell recognition by the host immune system and ultimately permits the manufacture of CAR-T cells from allogeneic T cells isolated from healthy donors [239]. For instance, TCR-deficient allogeneic T cells expressing anti-CD7 CAR could induce remarkable cytotoxicity against CD7-expressing leukemia and lymphoma cells in vivo without graft versus host disease (GvHD) occurrence [240]. Also, anti-CD19 CAR-T cells with depleted TCR and B2M did not provoke GVHD but retained antitumor responses in immunodeficient mice [241]. Thereby, CAR-positive TCR-negative T cells could be a reliable plan to establish next-generation CAR-T cells.

CAR-T cells with higher efficacy

The expression of immune checkpoints such as PD-1 and LAG3 and immunosuppressive biomolecules such as TGF-B in TME prevents significant and long-term activation of CAR-T cells in vivo. In 2019, Hu et al. showed that CRISPR/Cas9-mediated KO of PD-1 in anti-CD133 CAR-T cells resulted in potentiated proliferation and cytotoxicity in vitro and a murine glioma model [242]. In addition, PD-1-deficient anti-EGFRvIII CAR-T cells could stimulate a more efficient antitumor impact on EGFRvIII-positive glioblastoma cells with no adverse effect on T-cell phenotype or other biological activities [243]. Zhang et al. (2017) also showed that LAG-3 KO CAR-T cells exerted vigorous antigen-specific antitumor effect in a murine xenograft model of refractory B cell malignancy [244].

Given the existence of TGF-β in TME, many efforts have been made to establish TGFβ-receptors (R)-deficient CAR-T cells. Various reports have shown that CRISPR/Cas9-mediated genetic depletion of TGF-βRII causes upregulation of receptor tyrosine kinase-like orphan receptor 1 (ROR1) [245], B-cell maturation antigen (BCMA) [246], mesothelin [245], and PSMA [247], mediating specific CAR-T cell-induced antitumor activity by alleviating TGF-β.

CAR-T cells with minimized CRS occurrence

Due to its role in CRS development, granulocyte–macrophage colony-stimulating factor (GM-CSF) KO CAR-T cells have been suggested as a putative strategy to minimize CRS occurrence upon CAR-T cell administration [248]. GM-CSF is secreted at high levels by activated CAR-T cells and primarily contributes to activating monocytes and macrophages [249]. Sterner et al. (2019) displayed that GM-CSF KO CD19-specific CAR-T cells secreted lower GM-CSF in vivo, elicited more efficient antitumor activity, and improved OS in mice treated with GM-CSF-deficient CAR-T cells compared with mice treated with conventional CAR-T cells [250]. Preliminary clinical outcomes of one patient with non-Hodgkin’s lymphoma (NHL) and two patients with multiple myeloma (MM) treated with GM-CSF/TCR KO CAR-T cells demonstrated that CRISPR/Cas9-mediated ablation of GM-CSF/TCR had no adverse effect on CAR-T cell proliferation in these patients [251]. CRISPR-edited GM-CSF/TCR KO CAR-T cells exhibited marked persistence following administration and could reexpand following antigen exposure [251]. Noteworthy, all three patients treated with GM-CSF/TCR KO CAR-T cells attained complete response [251].

The off-target effect of CRISPR/Cas9

Although various CRISPR/Cas system classes have been developed, their wide-ranging application may be obstructed by various issues [252]. The main drawback of CRISPR/Cas9-driven gene editing is the correct prediction of its off-target function [253]. Off-target effects can be defined as accidental cleavage and mutations at untargeted genomic regions displaying a similar but not identical sequence compared with the target site. Indeed, a high incidence of off-target cleavages (≥ 50%) of RNA-guided endonuclease (RGEN)-stimulated mutations at sites other than the anticipated on-target site is the most eminent concern [254]. Another consideration for CRISPR/Cas9-directed gene editing is its editing efficiency [255]. For proficient gene editing treatment, efficient endonuclease accompanied by a dependable delivery system is paramount [256].

Various plans and methods have been designed and developed to improve the on-target effects and decrease possible off-target effects. Meanwhile, much effort has been invested in alleviating the off-target activity of CRISPR/Cas9 by creating multiple CRISPR/Cas systems that offer better fidelity and accuracy [257]. The genomic frameworks of the targeted DNA associated with the secondary structure of sgRNAs and their GC content (40–60% preferably) play a crucial role in defining the cleavage efficiency; the design of fitting sgRNAs with high on-target function using specific tools is urgently required [257]. In addition, the study of the cleavage potential of 218 sgRNAs using the in vitro mismatch cleavage assay or Surveyor assay signified that nucleotides at both PAM-distal and PAM-proximal site of the designed sgRNA are closely associated with the on-target efficiency [254]. For instance, G (but not C) is favored, and as the first base is closely neighboring the PAM, C (but not G) is favored at position 5 (the fifth base proximal to PAM) [258]. Furthermore, the distance between the PAM site and the start codon considerably varies the cleavage efficiency and target specificity. Also, adjusting the Cas9–sgRNA complex concentration by titrating the Cas9 and sgRNA delivery quantities is another approach that has been suggested to reduce off-target activity [258]. Of course, promoting specificity by decreasing the transfected DNA quantity may result in a decrease of on-target cleavage. The equilibrium between on-target cleavage effectiveness and off-target impacts thus has to be considered [258]. In addition, recent reports have delivered proof of concept that combinations of catalytically inactive Cas9 with endonuclease FokI nuclease domain (fCas9) could edit target DNA sequence with > 140-fold higher than wild-type Cas9 [259]. Further, wild-type Cas9 nuclease could be substituted with the D10 mutant nickase version of Cas9 and paired with two sgRNAs that cut only one strand. The paired nicking approach markedly decreases the off-target activity by 50–1500-fold in vitro [260]. During the last decade, researchers have concentrated on merging designer nuclease development [261], designing computational prediction programs and databases [262], and detecting high-throughput sequencing [263] to recognize off-target mutations and minimize off-target activity. Taken together, minimizing the off-target activity in the CRISPR/Cas9 system undeniably provides solid genotype–phenotype relations, thus enabling the realistic construction of gene editing statistics which, in turn, facilitates the clinical application of these CRISPR/Cas9 tools [258].

CRISPR screening

The development of CRISPR screening facilitates high-throughput probing of gene activities in multiple tumor biologies, such as tumor development, metastasis, synthetic lethal interrelation, therapeutic resistance, and response to immunotherapy, which are usually accomplished in vitro or in tumor-cell-bearing animals [264, 265]. CRISPR screening detects essential genes or genetic sequences that largely contribute to stimulating a particular action or phenotype for a cell type [266]. CRISPR/Cas9 exhibits better genetic editing ability, lower off-target effect, and more adaptability. It can be designed and carried out in various formats and affect either coding or noncoding regions in the genome compared with conventional approaches performed using RNAi or cDNA libraries [267]. Although the central idea of CRISPR screening is to knock out every gene (only one gene per cell) that could be significant (Fig. 4), the knockdown screen and activation screen are other types of CRISPR screening with a typical workflow. Firstly, designed sgRNAs are cloned into a lentivirus library and transduced into Cas9-expressing or dCas9-expressing cells at a low multiplicity of infection (MOI) to guarantee that only one copy of sgRNA is integrated per cell [268]. Secondly, CRISPR library-transduced cells undergo biology assay-based screening [268]. If the target gene changes cell fitness in the context of selection pressure, cells containing the sgRNA will be eliminated or potentiated among the population. Lastly, CRISPR screens leverage unique sgRNA sequences and next-generation sequencing (NGS) to detect alterations in sgRNA iteration following phenotypic selection [268]. As such technologies continue to advance, we believe that CRISPR screening will speed up investigations on the functional characterization of genetic materials and the discovery of new therapeutic targets.

Fig. 4
figure 4

CRISPR screening using pooled DNA oligos. In the most common types of CRISPR screening, a pool of oligos is designed to target a large number of genes. A library of lentiviruses is shaped from the oligos and applied to infect cells. CRISPR genome editing ablates several genes in various cells. Next-generation sequencing (NGS) is applied to define either present or absent genes. Importantly, genes for drug resistance or sensitivity can be detected. Meanwhile, negative screens define genes eliciting resistance, while positive screens define genes eliciting sensitivity

Conclusions and future directions

The CRISPR/Cas9 system allows one to edit a target sequence accurately in model organisms and humans for use in therapeutic analysis. Also, it is theoretically possible to treat infectious and genetic diseases and cancers. CRISPR/Cas9, as a customizable and easily applicable technique, facilitates the enlargement of complete genomic libraries for cancer patients.

Ongoing efforts are planned to maximize its specificity and thus tackle off-target cleavages. The recent progress in the CRISPR/Cas9 methodology reduces undesired mutations. Irrespective of minimizing the off-target action, which is a significant pitfall of gene editing tools, efficient delivery methods that promote their efficacy and constrain immune responses must be developed. Investigators are discovering diverse routes to fine-tune CRISPR delivery to specific cells in the human body. Cas9 ribonuclear proteins (RNPs) are now consistently exploited as a substitute for plasmid vectors for transporting the CRISPR reagent into target cells. This plan potentiates the efficiency, leads to a more transient Cas9 function, and will avert incorporation of vector sequences. Notwithstanding, this strategy does not constrain chromosomal rearrangements. As a final remark, it will be essential to optimize the efficacy, safety, and specificity of CRISPR/Cas9 before its clinical utility.

Availability of data and materials

Not applicable.

Abbreviations

CRISPR:

Clustered regularly interspaced short palindromic repeat

ZFN:

Zinc finger endonuclease

TALEN:

Transcription activator-like effector nuclease

CAR:

Chimeric antigen receptor

TGFβ:

Transforming growth factor beta

sgRNA:

Single-stranded guide RNA

crRNA:

CRISPR RNA

KO:

Knockout

PD1:

Programmed cell death protein 1

TCR:

T-cell receptor

GM-CSF:

Granulocyte–macrophage colony-stimulating factor

LAG3:

Lymphocyte activation gene 3

PAM:

Protospacer adjacent motif

DSBs:

DNA double-strand breaks

References

  1. Gupta SK, Shukla P. Gene editing for cell engineering: trends and applications. Crit Rev Biotechnol. 2017;37(5):672–84.

    Article  CAS  PubMed  Google Scholar 

  2. Dangi AK, Sinha R, Dwivedi S, Gupta SK, Shukla P. Cell line techniques and gene editing tools for antibody production: a review. Front Pharmacol. 2018;9:630.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Gaj T, Gersbach CA, Barbas CF III. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 2013;31(7):397–405.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Franic D, Dobrinic P, Korac P. Key achievements in gene therapy development and its promising progress with gene editing tools (ZFN, TALEN, CRISPR/CAS9). Mol Exp Biol Med. 2019;2(1):1–9.

    Article  Google Scholar 

  5. Kaboli S, Babazada H. CRISPR mediated genome engineering and its application in industry. Curr Issues Mol Biol. 2018;26(1):81–92.

    Article  PubMed  Google Scholar 

  6. Pattharaprachayakul N, Lee M, Incharoensakdi A, Woo HM. Current understanding of the cyanobacterial CRISPR-Cas systems and development of the synthetic CRISPR-Cas systems for cyanobacteria. Enzyme Microbial Technol. 2020;140:109619.

    Article  CAS  Google Scholar 

  7. Li R, Zatloukalova P, Muller P, Gil-Mir M, Kote S, Wilkinson S, Kemp AJ, Hernychova L, Wang Y, Ball KL. The MDM2 ligand Nutlin-3 differentially alters expression of the immune blockade receptors PD-L1 and CD276. Cell Mol Biol Lett. 2020;25(1):1–21.

    Article  CAS  Google Scholar 

  8. Mojica FJ, Rodriguez-Valera F. The discovery of CRISPR in archaea and bacteria. FEBS J. 2016;283(17):3162–9.

    Article  CAS  PubMed  Google Scholar 

  9. 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 

  10. Mojica FJ, Díez-Villaseñor C, García-Martínez J, Soria E. Intervening sequences of regularly spaced prokaryotic repeats derive from foreign genetic elements. J Mol Evol. 2005;60(2):174–82.

    Article  CAS  PubMed  Google Scholar 

  11. Horvath P, Barrangou R. CRISPR/Cas, the immune system of bacteria and archaea. Science. 2010;327(5962):167–70.

    Article  CAS  PubMed  Google Scholar 

  12. Zhang M, Eshraghian EA, Al Jammal O, Zhang Z, Zhu X. CRISPR technology: the engine that drives cancer therapy. Biomed Pharmacother. 2021;133:111007.

    Article  CAS  PubMed  Google Scholar 

  13. Chen M, Mao A, Xu M, Weng Q, Mao J, Ji J. CRISPR-Cas9 for cancer therapy: opportunities and challenges. Cancer Lett. 2019;447:48–55.

    Article  CAS  PubMed  Google Scholar 

  14. Zhang H, Qin C, An C, Zheng X, Wen S, Chen W, Liu X, Lv Z, Yang P, Xu W. Application of the CRISPR/Cas9-based gene editing technique in basic research, diagnosis, and therapy of cancer. Mol Cancer. 2021;20(1):1–22.

    Article  Google Scholar 

  15. Rodríguez TC, Dadafarin S, Pratt HE, Liu P, Amrani N, Zhu LJ. Genome-wide detection and analysis of CRISPR-Cas off-targets. Progress in molecular biology and translational science. 181: Elsevier; 2021. p. 31–43.

  16. Cheng DK, Oni TE, Thalappillil JS, Park Y, Ting H-C, Alagesan B, Prasad NV, Addison K, Rivera KD, Pappin DJ. Oncogenic KRAS engages an RSK1/NF1 pathway to inhibit wild-type RAS signaling in pancreatic cancer. Proc Natl Acad Sci. 2021;118(21).

  17. Dufva O, Koski J, Maliniemi P, Ianevski A, Klievink J, Leitner J, Pölönen P, Hohtari H, Saeed K, Hannunen T. Integrated drug profiling and CRISPR screening identify essential pathways for CAR T-cell cytotoxicity. Blood. 2020;135(9):597–609.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Hu W, Zi Z, Jin Y, Li G, Shao K, Cai Q, Ma X, Wei F. CRISPR/Cas9-mediated PD-1 disruption enhances human mesothelin-targeted CAR T cell effector functions. Cancer Immunol Immunother. 2019;68(3):365–77.

    Article  CAS  PubMed  Google Scholar 

  19. Iyer V, Boroviak K, Thomas M, Doe B, Riva L, Ryder E, Adams DJ. No unexpected CRISPR-Cas9 off-target activity revealed by trio sequencing of gene-edited mice. PLoS Genet. 2018;14(7):e1007503.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Razeghian E, Nasution MKM, Rahman HS, Gardanova ZR, Abdelbasset WK, Aravindhan S, Bokov DO, Suksatan W, Nakhaei P, Shariatzadeh S, et al. A deep insight into CRISPR/Cas9 application in CAR-T cell-based tumor immunotherapies. Stem Cell Res Ther. 2021;12(1):428.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Liang X, Potter J, Kumar S, Zou Y, Quintanilla R, Sridharan M, Carte J, Chen W, Roark N, Ranganathan S. Rapid and highly efficient mammalian cell engineering via Cas9 protein transfection. J Biotechnol. 2015;208:44–53.

    Article  CAS  PubMed  Google Scholar 

  22. Liu M, Han X, Liu H, Chen D, Li Y, Hu W. The effects of CRISPR-Cas9 knockout of the TGF-β1 gene on antler cartilage cells in vitro. Cell Mol Biol Lett. 2019;24(1):1–12.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Bhaya D, Davison M, Barrangou R. CRISPR-Cas systems in bacteria and archaea: versatile small RNAs for adaptive defense and regulation. Annu Rev Genet. 2011;45:273–97.

    Article  CAS  PubMed  Google Scholar 

  24. 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 

  25. Faure G, Shmakov SA, Makarova KS, Wolf YI, Crawley AB, Barrangou R, Koonin EV. Comparative genomics and evolution of trans-activating RNAs in class 2 CRISPR-Cas systems. RNA Biol. 2019;16(4):435–48.

    Article  PubMed  Google Scholar 

  26. Nishimasu H, Ran FA, Hsu PD, Konermann S, Shehata SI, Dohmae N, Ishitani R, Zhang F, Nureki O. 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 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Hsu PD, Lander ES, Zhang F. Development and applications of CRISPR-Cas9 for genome engineering. Cell. 2014;157(6):1262–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Allen D, Rosenberg M, Hendel A. Using synthetically engineered guide RNAs to enhance CRISPR genome editing systems in mammalian cells. Front Genome Editing. 2021;2.

  30. Pourcel C, Salvignol G, Vergnaud G. CRISPR elements in Yersinia pestis acquire new repeats by preferential uptake of bacteriophage DNA, and provide additional tools for evolutionary studies. Microbiology (Reading). 2005;151(Pt 3):653–63.

    Article  CAS  Google Scholar 

  31. Deveau H, Barrangou R, Garneau JE, Labonté J, Fremaux C, Boyaval P, Romero DA, Horvath P, Moineau S. Phage response to CRISPR-encoded resistance in Streptococcus thermophilus. J Bacteriol. 2008;190(4):1390–400.

    Article  CAS  PubMed  Google Scholar 

  32. Zhang F, Wen Y, Guo X. CRISPR/Cas9 for genome editing: progress, implications and challenges. Hum Mol Genet. 2014;23(R1):R40–6.

    Article  CAS  PubMed  Google Scholar 

  33. Hryhorowicz M, Lipiński D, Zeyland J, Słomski R. CRISPR/Cas9 immune system as a tool for genome engineering. Arch Immunol Ther Exp (Warsz). 2017;65(3):233–40.

    Article  CAS  Google Scholar 

  34. Shojaei Baghini S, Gardanova ZR, Zekiy AO, Shomali N, Tosan F, Jarahian M. Optimizing sgRNA to improve CRISPR/Cas9 knockout efficiency: special focus on human and animal cell. Front Bioeng Biotechnol 2021;9:775309.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Koonin EV, Makarova KS, Zhang F. Diversity, classification and evolution of CRISPR-Cas systems. Curr Opin Microbiol. 2017;37:67–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Makarova KS, Haft DH, Barrangou R, Brouns SJ, Charpentier E, Horvath P, Moineau S, Mojica FJ, Wolf YI, Yakunin AF. Evolution and classification of the CRISPR–Cas systems. Nat Rev Microbiol. 2011;9(6):467–77.

    Article  CAS  PubMed  Google Scholar 

  37. Yamano T, Nishimasu H, Zetsche B, Hirano H, Slaymaker IM, Li Y, Fedorova I, Nakane T, Makarova KS, Koonin EV, et al. Crystal structure of Cpf1 in complex with guide RNA and target DNA. Cell. 2016;165(4):949–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Hou Z, Zhang Y, Propson NE, Howden SE, Chu LF, Sontheimer EJ, Thomson JA. Efficient genome engineering in human pluripotent stem cells using Cas9 from Neisseria meningitidis. Proc Natl Acad Sci U S A. 2013;110(39):15644–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Zhou J, Shen B, Zhang W, Wang J, Yang J, Chen L, Zhang N, Zhu K, Xu J, Hu B. One-step generation of different immunodeficient mice with multiple gene modifications by CRISPR/Cas9 mediated genome engineering. Int J Biochem Cell Biol. 2014;46:49–55.

    Article  CAS  PubMed  Google Scholar 

  40. Kennedy EM, Kornepati AV, Mefferd AL, Marshall JB, Tsai K, Bogerd HP, Cullen BR. Optimization of a multiplex CRISPR/Cas system for use as an antiviral therapeutic. Methods. 2015;91:82–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Uranga M, Aragonés V, Selma S, Vázquez-Vilar M, Orzáez D, Daròs JA. Efficient Cas9 multiplex editing using unspaced sgRNA arrays engineering in a Potato virus X vector. Plant J. 2021;106(2):555–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Wang J, Yang J, Li D, Li J. Technologies for targeting DNA methylation modifications: basic mechanism and potential application in cancer. Biochim Biophys Acta Rev Cancer 2021;1875(1):188454.

    Article  CAS  PubMed  Google Scholar 

  43. Wu X, Tay JK, Goh CK, Chan C, Lee YH, Springs SL, Loh KS, Lu TK, Yu H. Digital CRISPR-based method for the rapid detection and absolute quantification of nucleic acids. Biomaterials. 2021;274:120876.

    Article  CAS  PubMed  Google Scholar 

  44. Wei J, Alfajaro MM, DeWeirdt PC, Hanna RE, Lu-Culligan WJ, Cai WL, Strine MS, Zhang S-M, Graziano VR, Schmitz CO. Genome-wide CRISPR screens reveal host factors critical for SARS-CoV-2 infection. Cell. 2021;184(1):76–91.

    Article  CAS  PubMed  Google Scholar 

  45. Zhao K-R, Wang L, Liu P-F, Hang X-M, Wang H-Y, Ye S-Y, Liu Z-J, Liang G-X. A signal-switchable electrochemiluminescence biosensor based on the integration of spherical nucleic acid and CRISPR/Cas12a for multiplex detection of HIV/HPV DNAs. Sens Actuators B: Chem. 2021;346:130485.

    Article  CAS  Google Scholar 

  46. Kennedy EM, Cullen BR. Bacterial CRISPR/Cas DNA endonucleases: a revolutionary technology that could dramatically impact viral research and treatment. Virology. 2015;479:213–20.

    Article  PubMed  CAS  Google Scholar 

  47. Lin H, Li G, Peng X, Deng A, Ye L, Shi L, Wang T, He J. The use of CRISPR/Cas9 as a tool to study human infectious viruses. Front Cell Infect Microbiol. 2021;11.

  48. Saayman S, Ali SA, Morris KV, Weinberg MS. The therapeutic application of CRISPR/Cas9 technologies for HIV. Expert Opin Biol Ther. 2015;15(6):819–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Azar M, Aghazadeh H, Mohammed HN, Sara MRS, Hosseini A, Shomali N, Tamjidifar R, Tarzi S, Mansouri M, Sarand SP, et al. miR-193a-5p as a promising therapeutic candidate in colorectal cancer by reducing 5-FU and Oxaliplatin chemoresistance by targeting CXCR4. Int Immunopharmacol. 2021;92:107355.

    Article  CAS  PubMed  Google Scholar 

  50. Ebina H, Misawa N, Kanemura Y, Koyanagi Y. Harnessing the CRISPR/Cas9 system to disrupt latent HIV-1 provirus. Sci Rep. 2013;3(1):1–7.

    Article  Google Scholar 

  51. Hou P, Chen S, Wang S, Yu X, Chen Y, Jiang M, Zhuang K, Ho W, Hou W, Huang J, et al. Genome editing of CXCR4 by CRISPR/cas9 confers cells resistant to HIV-1 infection. Sci Rep. 2015;5:15577.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. De Paoli P, Carbone A. Carcinogenic viruses and solid cancers without sufficient evidence of causal association. Int J Cancer. 2013;133(7):1517–29.

    Article  PubMed  CAS  Google Scholar 

  53. Donà S, Borsetto D, Fussey J, Biscaro V, Vian E, Spinato G, Menegaldo A, Da Mosto MC, Rigoli R, Polesel J. Association between hepatitis C and B viruses and head and neck squamous cell carcinoma. J Clin Virol. 2019;121:104209.

    Article  PubMed  CAS  Google Scholar 

  54. Zhen S, Hua L, Liu YH, Gao LC, Fu J, Wan DY, Dong LH, Song HF, Gao X. Harnessing the clustered regularly interspaced short palindromic repeat (CRISPR)/CRISPR-associated Cas9 system to disrupt the hepatitis B virus. Gene Ther. 2015;22(5):404–12.

    Article  CAS  PubMed  Google Scholar 

  55. Dong C, Qu L, Wang H, Wei L, Dong Y, Xiong S. Targeting hepatitis B virus cccDNA by CRISPR/Cas9 nuclease efficiently inhibits viral replication. Antiviral Res. 2015;118:110–7.

    Article  CAS  PubMed  Google Scholar 

  56. Kostyushev D, Kostyusheva A, Brezgin S, Zarifyan D, Utkina A, Goptar I, Chulanov V. Suppressing the NHEJ pathway by DNA-PKcs inhibitor NU7026 prevents degradation of HBV cccDNA cleaved by CRISPR/Cas9. Sci Rep. 2019;9(1):1847.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Chen Y, Jiang H, Wang T, He D, Tian R, Cui Z, Tian X, Gao Q, Ma X, Yang J, et al. In vitro and in vivo growth inhibition of human cervical cancer cells via human papillomavirus E6/E7 mRNAs’ cleavage by CRISPR/Cas13a system. Antiviral Res. 2020;178:104794.

    Article  CAS  PubMed  Google Scholar 

  58. Ling K, Yang L, Yang N, Chen M, Wang Y, Liang S, Li Y, Jiang L, Yan P, Liang Z. Gene targeting of HPV18 E6 and E7 synchronously by nonviral transfection of CRISPR/Cas9 system in cervical cancer. Hum Gene Ther. 2020;31(5–6):297–308.

    Article  CAS  PubMed  Google Scholar 

  59. Hu Z, Yu L, Zhu D, Ding W, Wang X, Zhang C, Wang L, Jiang X, Shen H, He D, et al. Disruption of HPV16-E7 by CRISPR/Cas system induces apoptosis and growth inhibition in HPV16 positive human cervical cancer cells. Biomed Res Int. 2014;2014:612823.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Akasu M, Shimada S, Kabashima A, Akiyama Y, Shimokawa M, Akahoshi K, Kudo A, Yamaoka S, Tanabe M, Tanaka S. Intrinsic activation of β-catenin signaling by CRISPR/Cas9-mediated exon skipping contributes to immune evasion in hepatocellular carcinoma. 2021.

  61. Srour N, Villarreal OD, Yu Z, Preston S, Miller WH, Szewczyk MM, Barsyte-Lovejoy D, Xu H, del Rincón SV, Richard S. PRMT7 ablation stimulates anti-tumor immunity and sensitizes melanoma to immune checkpoint blockade. bioRxiv. 2021.

  62. Shomali N, Hatamnezhad LS, Tarzi S, Tamjidifar R, Xu H, Shotorbani SS. Heat shock proteins regulating toll-like receptors and the immune system could be a novel therapeutic target for melanoma. Curr Mol Med. 2021;21(1):15–24.

    Article  CAS  PubMed  Google Scholar 

  63. Zhen S, Lu J, Liu Y-H, Chen W, Li X. Synergistic antitumor effect on cervical cancer by rational combination of PD1 blockade and CRISPR-Cas9-mediated HPV knockout. Cancer Gene Ther. 2020;27(3):168–78.

    Article  CAS  PubMed  Google Scholar 

  64. Hosseinzadeh R, Feizisani F, Shomali N, Abdelbasset WK, Hemmatzadeh M, Gholizadeh Navashenaq J, Jadidi-Niaragh F, Bokov DO, Janebifam M, Mohammadi H. PD-1/PD-L1 blockade: prospectives for immunotherapy in cancer and autoimmunity. IUBMB Life. 2021;73(11):1293–306.

    Article  CAS  PubMed  Google Scholar 

  65. Bikard D, Barrangou R. Using CRISPR-Cas systems as antimicrobials. Curr Opin Microbiol. 2017;37:155–60.

    Article  CAS  PubMed  Google Scholar 

  66. Strich JR, Chertow DS. CRISPR-Cas biology and its application to infectious diseases. J Clin Microbiol. 2019;57(4):e01307-e1318.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Milunsky A, Milunsky JM. Genetic disorders and the fetus: diagnosis, prevention, and treatment. Hoboken: Wiley; 2015.

    Book  Google Scholar 

  68. Kaplanis J, Samocha KE, Wiel L, Zhang Z, Arvai KJ, Eberhardt RY, Gallone G, Lelieveld SH, Martin HC, McRae JF. Evidence for 28 genetic disorders discovered by combining healthcare and research data. Nature. 2020;586(7831):757–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Xu X, Wan T, Xin H, Li D, Pan H, Wu J, Ping Y. Delivery of CRISPR/Cas9 for therapeutic genome editing. J Gene Med. 2019;21(7):e3107.

    Article  PubMed  CAS  Google Scholar 

  70. Sharma G, Sharma AR, Bhattacharya M, Lee S-S, Chakraborty C. CRISPR-Cas9: a preclinical and clinical perspective for the treatment of human diseases. Mol Ther. 2021;29(2):571–86.

    Article  CAS  PubMed  Google Scholar 

  71. Asher DR, Thapa K, Dharia SD, Khan N, Potter RA, Rodino-Klapac LR, Mendell JR. Clinical development on the frontier: gene therapy for Duchenne muscular dystrophy. Expert Opin Biol Ther. 2020;20(3):263–74.

    Article  CAS  PubMed  Google Scholar 

  72. Izmiryan A, Ganier C, Bovolenta M, Schmitt A, Mavilio F, Hovnanian A. Ex vivo COL7A1 correction for recessive dystrophic epidermolysis bullosa using CRISPR/Cas9 and homology-directed repair. Mol Therapy-Nucleic Acids. 2018;12:554–67.

    Article  CAS  Google Scholar 

  73. Gurumurthy CB, Sato M, Nakamura A, Inui M, Kawano N, Islam MA, Ogiwara S, Takabayashi S, Matsuyama M, Nakagawa S. Creation of CRISPR-based germline-genome-engineered mice without ex vivo handling of zygotes by i-GONAD. Nat Protoc. 2019;14(8):2452–82.

    Article  CAS  PubMed  Google Scholar 

  74. Friedmann T, Roblin R. Gene therapy for human genetic disease? Science. 1972;175(4025):949–55.

    Article  CAS  PubMed  Google Scholar 

  75. Dowaidar M. Genome-wide association studies (GWAS) have revolutionized our view of human health and disease genetics and offered novel gene therapy targets. 2021.

  76. Pavani G, Fabiano A, Laurent M, Amor F, Cantelli E, Chalumeau A, Maule G, Tachtsidi A, Concordet J-P, Cereseto A. Correction of β-thalassemia by CRISPR/Cas9 editing of the α-globin locus in human hematopoietic stem cells. Blood Adv. 2021;5(5):1137–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Frangoul H, Bobruff Y, Cappellini MD, Corbacioglu S, Fernandez CM, De la Fuente J, Grupp SA, Handgretinger R, Ho TW, Imren S. Safety and efficacy of CTX001 in patients with transfusion-dependent β-thalassemia and sickle cell disease: early results from the climb THAL-111 and climb SCD-121 studies of autologous CRISPR-CAS9-modified CD34+ hematopoietic stem and progenitor cells. Blood. 2020;136:3–4.

    Article  Google Scholar 

  78. Xie F, Ye L, Chang JC, Beyer AI, Wang J, Muench MO, Kan YW. Seamless gene correction of β-thalassemia mutations in patient-specific iPSCs using CRISPR/Cas9 and piggyBac. Genome Res. 2014;24(9):1526–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Li N, Gou S, Wang J, Zhang Q, Huang X, Xie J, Li L, Jin Q, Ouyang Z, Chen F. CRISPR/Cas9-mediated gene correction in newborn rabbits with hereditary tyrosinemia type I. Mol Ther. 2021;29(3):1001–15.

    Article  CAS  PubMed  Google Scholar 

  80. Min Y-L, Bassel-Duby R, Olson EN. CRISPR correction of Duchenne muscular dystrophy. Annu Rev Med. 2019;70:239–55.

    Article  CAS  PubMed  Google Scholar 

  81. Young CS, Hicks MR, Ermolova NV, Nakano H, Jan M, Younesi S, Karumbayaram S, Kumagai-Cresse C, Wang D, Zack JA. A single CRISPR-Cas9 deletion strategy that targets the majority of DMD patients restores dystrophin function in hiPSC-derived muscle cells. Cell Stem Cell. 2016;18(4):533–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Park C-Y, Kim DH, Son JS, Sung JJ, Lee J, Bae S, Kim J-H, Kim D-W, Kim J-S. Functional correction of large factor VIII gene chromosomal inversions in hemophilia A patient-derived iPSCs using CRISPR-Cas9. Cell Stem Cell. 2015;17(2):213–20.

    Article  CAS  PubMed  Google Scholar 

  83. Guan Y, Ma Y, Li Q, Sun Z, Ma L, Wu L, Wang L, Zeng L, Shao Y, Chen Y. CRISPR/Cas9-mediated somatic correction of a novel coagulator factor IX gene mutation ameliorates hemophilia in mouse. EMBO Mol Med. 2016;8(5):477–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Marangi M, Pistritto G. Innovative therapeutic strategies for cystic fibrosis: moving forward to CRISPR technique. Front Pharmacol. 2018;9:396.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  85. Tu Z, Yang W, Yan S, Guo X, Li X-J. CRISPR/Cas9: a powerful genetic engineering tool for establishing large animal models of neurodegenerative diseases. Mol Neurodegener. 2015;10(1):1–8.

    Article  CAS  Google Scholar 

  86. Rahman S, Datta M, Kim J, Jan AT, editors. CRISPR/Cas: An intriguing genomic editing tool with prospects in treating neurodegenerative diseases. Seminars in cell & developmental biology; 2019: Elsevier.

  87. Qian Y, Zhao D, Sui T, Chen M, Liu Z, Liu H, Zhang T, Chen S, Lai L, Li Z. Efficient and precise generation of Tay-Sachs disease model in rabbit by prime editing system. Cell Discovery. 2021;7(1):1–3.

    Article  CAS  Google Scholar 

  88. Lee B, Lee K, Panda S, Gonzales-Rojas R, Chong A, Bugay V, Park HM, Brenner R, Murthy N, Lee HY. Nanoparticle delivery of CRISPR into the brain rescues a mouse model of fragile X syndrome from exaggerated repetitive behaviours. Nat Biomed Eng. 2018;2(7):497–507.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Yrigollen CM, Davidson BL. CRISPR to the rescue: advances in gene editing for the FMR1 gene. Brain Sci. 2019;9(1):17.

    Article  CAS  PubMed Central  Google Scholar 

  90. Zhang B. CRISPR/Cas gene therapy. J Cell Physiol. 2021;236(4):2459–81.

    Article  CAS  PubMed  Google Scholar 

  91. Wan T, Ping Y. Delivery of genome-editing biomacromolecules for treatment of lung genetic disorders. Adv Drug Deliv Rev. 2021;168:196–216.

    Article  CAS  PubMed  Google Scholar 

  92. Lockyer EJ. The potential of CRISPR-Cas9 for treating genetic disorders. Biosci Horizons Int J Stud Res. 2016;9.

  93. Luthra R, Kaur S, Bhandari K. Applications of CRISPR as a potential therapeutic. Life Sci. 2021;284:119908.

    Article  CAS  PubMed  Google Scholar 

  94. Pandey V, Tripathi A, Bhushan R, Ali A, Dubey P, Therapy G. Application of CRISPR/Cas9 genome editing in genetic disorders: a systematic review up to date. J Genet Syndr Gene Ther. 2017;8(2):1–10.

    Article  Google Scholar 

  95. Li C, Mei H, Hu Y. Applications and explorations of CRISPR/Cas9 in CAR T-cell therapy. Brief Funct Genomics. 2020;19(3):175–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Yang Y, Chen X, Yao W, Cui X, Li N, Lin Z, Zhao B, Miao J. Esterase D stabilizes FKBP25 to suppress mTORC1. Cell Mol Biol Lett. 2021;26(1):50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Li W, Wu L, Jia H, Lin Z, Zhong R, Li Y, Jiang C, Liu S, Zhou X, Zhang E. The low-complexity domains of the KMT2D protein regulate histone monomethylation transcription to facilitate pancreatic cancer progression. Cell Mol Biol Lett. 2021;26(1):45.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  98. Zhang S, Zhang F, Chen Q, Wan C, Xiong J, Xu J. CRISPR/Cas9-mediated knockout of NSD1 suppresses the hepatocellular carcinoma development via the NSD1/H3/Wnt10b signaling pathway. J Exp Clin Cancer Res. 2019;38(1):467.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  99. Guan L, Zhu S, Han Y, Yang C, Liu Y, Qiao L, Li X, Li H, Lin J. Knockout of CTNNB1 by CRISPR-Cas9 technology inhibits cell proliferation through the Wnt/β-catenin signaling pathway. Biotechnol Lett. 2018;40(3):501–8.

    Article  CAS  PubMed  Google Scholar 

  100. Jin C, Yuan FL, Gu YL, Li X, Liu MF, Shen XM, Liu B, Zhu MQ. Over-expression of ASIC1a promotes proliferation via activation of the β-catenin/LEF-TCF axis and is associated with disease outcome in liver cancer. Oncotarget. 2017;8(16):25977–88.

    Article  PubMed  Google Scholar 

  101. Novak A, Dedhar S. Signaling through beta-catenin and Lef/Tcf. Cell Mol Life Sci. 1999;56(5–6):523–37.

    Article  CAS  PubMed  Google Scholar 

  102. Liu Y, Zhuang H, Cao F, Li J, Guo Y, Zhang J, Zhao Q, Liu Y. Shc3 promotes hepatocellular carcinoma stemness and drug resistance by interacting with β-catenin to inhibit its ubiquitin degradation pathway. Cell Death Dis. 2021;12(3):1–15.

    Article  Google Scholar 

  103. Chen T, Lin J, Tang D, Zhang M, Wen F, Xue D, Zhang H. Paris saponin H suppresses human hepatocellular carcinoma (HCC) by inactivation of Wnt/β-catenin pathway in vitro and in vivo. Int J Clin Exp Pathol. 2019;12(8):2875.

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Fawzy IO, Hamza MT, Hosny KA, Esmat G, Abdelaziz AI. Abrogating the interplay between IGF2BP1, 2 and 3 and IGF1R by let-7i arrests hepatocellular carcinoma growth. Growth Factors. 2016;34(1–2):42–50.

    Article  CAS  PubMed  Google Scholar 

  105. Cai X, Chen Y, Man D, Yang B, Feng X, Zhang D, Chen J, Wu J. RBM15 promotes hepatocellular carcinoma progression by regulating N6-methyladenosine modification of YES1 mRNA in an IGF2BP1-dependent manner. Cell Death Discovery. 2021;7(1):315.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Zhang J, Hu K, Yang YQ, Wang Y, Zheng YF, Jin Y, Li P, Cheng L. LIN28B-AS1-IGF2BP1 binding promotes hepatocellular carcinoma cell progression. Cell Death Dis. 2020;11(9):741.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  107. Müller S, Glaß M, Singh AK, Haase J, Bley N, Fuchs T, Lederer M, Dahl A, Huang H, Chen J, et al. IGF2BP1 promotes SRF-dependent transcription in cancer in a m6A- and miRNA-dependent manner. Nucleic Acids Res. 2019;47(1):375–90.

    Article  PubMed  CAS  Google Scholar 

  108. Joseph ELM, Laheurte C, Jary M, Boullerot L, Asgarov K, Gravelin E, Bouard A, Rangan L, Dosset M, Borg C. Immunoregulation and clinical implications of ANGPT2/TIE2+ M-MDSC signature in non–small cell lung cancer. Cancer Immunol Res. 2020;8(2):268–79.

    Article  CAS  Google Scholar 

  109. Xie J-y, Wei J-x, Lv L-h, Han Q-f, Yang W-b, Li G-l, Wang P-x, Wu S-b, Duan J-x, Zhuo W-f. Angiopoietin-2 induces angiogenesis via exosomes in human hepatocellular carcinoma. Cell Commun Signaling. 2020;18(1):1–13.

    Article  CAS  Google Scholar 

  110. Huang H, Bhat A, Woodnutt G, Lappe R. Targeting the ANGPT–TIE2 pathway in malignancy. Nat Rev Cancer. 2010;10(8):575–85.

    Article  CAS  PubMed  Google Scholar 

  111. Xu J, Liu F, Xia Z, He K, Xiang G. MiR-3188 regulates the proliferation and apoptosis of hepatocellular carcinoma cells by targeting CXCL14. Biomark Med. 2021;15(17):1611–21.

    CAS  PubMed  Google Scholar 

  112. Zhu H-R, Huang R-Z, Yu X-N, Shi X, Bilegsaikhan E, Guo H-Y, Song G-Q, Weng S-Q, Dong L, Janssen HL. Microarray expression profiling of microRNAs reveals potential biomarkers for hepatocellular carcinoma. Tohoku J Exp Med. 2018;245(2):89–98.

    Article  CAS  PubMed  Google Scholar 

  113. Zhou S-j, Deng Y-l, Liang H-f, Jaoude JC, Liu F-y. Hepatitis B virus X protein promotes CREB-mediated activation of miR-3188 and Notch signaling in hepatocellular carcinoma. Cell Death Diff. 2017;24(9):1577–87.

    Article  CAS  Google Scholar 

  114. Wei L, Lee D, Law C-T, Zhang MS, Shen J, Chin DW-C, Zhang A, Tsang FH-C, Wong CL-S, Ng IO-L, et al. Genome-wide CRISPR/Cas9 library screening identified PHGDH as a critical driver for Sorafenib resistance in HCC. Nat Commun. 2019;10(1):4681.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  115. Zare K, Shademan M, Ghahramani Seno MM, Dehghani H. CRISPR/Cas9 knockout strategies to ablate CCAT1 lncRNA gene in cancer cells. Biol Proced Online. 2018;20:21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Shang A, Wang W, Gu C, Chen W, Lu W, Sun Z, Li D. Long non-coding RNA CCAT1 promotes colorectal cancer progression by regulating miR-181a-5p expression. Aging (Albany NY). 2020;12(9):8301–20.

    Article  CAS  Google Scholar 

  117. Takeda H, Kataoka S, Nakayama M, Ali MAE, Oshima H, Yamamoto D, Park JW, Takegami Y, An T, Jenkins NA, et al. CRISPR-Cas9-mediated gene knockout in intestinal tumor organoids provides functional validation for colorectal cancer driver genes. Proc Natl Acad Sci U S A. 2019;116(31):15635–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Ganguly K, Krishn SR, Rachagani S, Jahan R, Shah A, Nallasamy P, Rauth S, Atri P, Cox JL, Pothuraju R, et al. Secretory mucin 5AC promotes neoplastic progression by augmenting KLF4-mediated pancreatic cancer cell stemness. Cancer Res. 2021;81(1):91–102.

    Article  CAS  PubMed  Google Scholar 

  119. Kato S, Hokari R, Crawley S, Gum J, Ahn DH, Kim JW, Kwon SW, Miura S, Basbaum CB, Kim YS. MUC5AC mucin gene regulation in pancreatic cancer cells. Int J Oncol. 2006;29(1):33–40.

    CAS  PubMed  Google Scholar 

  120. Kaprio T, Hagström J, Mustonen H, Koskensalo S, Andersson LC, Haglund C. REG4 independently predicts better prognosis in non-mucinous colorectal cancer. PLoS ONE. 2014;9(10):e109600.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  121. Pothuraju R, Rachagani S, Krishn SR, Chaudhary S, Nimmakayala RK, Siddiqui JA, Ganguly K, Lakshmanan I, Cox JL, Mallya K, et al. Molecular implications of MUC5AC-CD44 axis in colorectal cancer progression and chemoresistance. Mol Cancer. 2020;19(1):37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. O’Cathail SM, Wu CH, Thomas R, Hawkins MA, Maughan TS, Lewis A. NRF2 mediates therapeutic resistance to chemoradiation in colorectal cancer through a metabolic switch. Antioxidants (Basel) 2021;10(9):1380.

    Article  CAS  Google Scholar 

  123. Li T, Liu D, Lei X, Jiang Q. Par3L enhances colorectal cancer cell survival by inhibiting Lkb1/AMPK signaling pathway. Biochem Biophys Res Commun. 2017;482(4):1037–41.

    Article  CAS  PubMed  Google Scholar 

  124. Hu X, Zhang L, Li Y, Ma X, Dai W, Gao X, Rao X, Fu G, Wang R, Pan M, et al. Organoid modelling identifies that DACH1 functions as a tumour promoter in colorectal cancer by modulating BMP signalling. EBioMedicine. 2020;56:102800.

    Article  PubMed  PubMed Central  Google Scholar 

  125. Xu H, Yu S, Yuan X, Xiong J, Kuang D, Pestell RG, Wu K. DACH1 suppresses breast cancer as a negative regulator of CD44. Sci Rep. 2017;7(1):4361.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  126. Kumagai T, Tomari K, Shimizu T, Takeda K. Alteration of gene expression in response to bone morphogenetic protein-2 in androgen-dependent human prostate cancer LNCaP cells. Int J Mol Med. 2006;17(2):285–91.

    CAS  PubMed  Google Scholar 

  127. Yan W, Wu K, Herman JG, Brock MV, Fuks F, Yang L, Zhu H, Li Y, Yang Y, Guo M. Epigenetic regulation of DACH1, a novel Wnt signaling component in colorectal cancer. Epigenetics. 2013;8(12):1373–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Ramírez-Ramírez R, Gutiérrez-Angulo M, Peregrina-Sandoval J, Moreno-Ortiz JM, Franco-Topete RA, Cerda-Camacho FJ, Ayala-Madrigal ML. Somatic deletion of KDM1A/LSD1 gene is associated to advanced colorectal cancer stages. J Clin Pathol. 2020;73(2):107–11.

    Article  PubMed  CAS  Google Scholar 

  129. Miller SA, Policastro RA, Savant SS, Sriramkumar S, Ding N, Lu X, Mohammad HP, Cao S, Kalin JH, Cole PA, et al. Lysine-specific demethylase 1 mediates AKT activity and promotes epithelial-to-mesenchymal transition in PIK3CA-mutant colorectal cancer. Mol Cancer Res. 2020;18(2):264–77.

    Article  CAS  PubMed  Google Scholar 

  130. Deb G, Wingelhofer B, Amaral FMR, Maiques-Diaz A, Chadwick JA, Spencer GJ, Williams EL, Leong HS, Maes T, Somervaille TCP. Pre-clinical activity of combined LSD1 and mTORC1 inhibition in MLL-translocated acute myeloid leukaemia. Leukemia. 2020;34(5):1266–77.

    Article  CAS  PubMed  Google Scholar 

  131. Gao J, Ren SR, Wang BR, Guo Q, Feng TT, Wang D, Liu JH, Tong JY, Shi LH. Knockout of LSD1 gene by CRISPR/Cas9 system significantly inhibited proliferation and expression of CD235a in K562 cells. Zhongguo Shi Yan Xue Ye Xue Za Zhi. 2017;25(5):1327–33.

    PubMed  Google Scholar 

  132. Park DE, Cheng J, McGrath JP, Lim MY, Cushman C, Swanson SK, Tillgren ML, Paulo JA, Gokhale PC, Florens L, et al. Merkel cell polyomavirus activates LSD1-mediated blockade of non-canonical BAF to regulate transformation and tumorigenesis. Nat Cell Biol. 2020;22(5):603–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Kim S, Bolatkan A, Kaneko S, Ikawa N, Asada K, Komatsu M, Hayami S, Ojima H, Abe N, Yamaue H, et al. Deregulation of the histone lysine-specific demethylase 1 is involved in human hepatocellular carcinoma. Biomolecules. 2019;9(12):810.

    Article  CAS  PubMed Central  Google Scholar 

  134. Faryal R. Role of miRNAs in breast cancer. Asian Pac J Cancer Prev. 2011;12:3175–80.

    PubMed  Google Scholar 

  135. Zare M, Bastami M, Solali S, Alivand MR. Aberrant miRNA promoter methylation and EMT-involving miRNAs in breast cancer metastasis: diagnosis and therapeutic implications. J Cell Physiol. 2018;233(5):3729–44.

    Article  CAS  PubMed  Google Scholar 

  136. Hannafon BN, Cai A, Calloway CL, Xu Y-F, Zhang R, Fung K-M, Ding W-Q. miR-23b and miR-27b are oncogenic microRNAs in breast cancer: evidence from a CRISPR/Cas9 deletion study. BMC Cancer. 2019;19(1):1–12.

    Article  CAS  Google Scholar 

  137. Dai Y-H, Wang Y-F, Shen P-C, Lo C-H, Yang J-F, Lin C-S, Chao H-L, Huang W-Y. Gene-associated methylation status of ST14 as a predictor of survival and hormone receptor positivity in breast Cancer. BMC Cancer. 2021;21(1):1–14.

    Article  Google Scholar 

  138. Kim KY, Yoon M, Cho Y, Lee K-H, Park S, Lee S-r, Choi S-Y, Lee D, Yang C, Cho EH. Targeting metastatic breast cancer with peptide epitopes derived from autocatalytic loop of Prss14/ST14 membrane serine protease and with monoclonal antibodies. J Exp Clin Cancer Res. 2019;38(1):1–17.

    Article  Google Scholar 

  139. Zhang J, Song Y, Shi Q, Fu L. Research progress on FASN and MGLL in the regulation of abnormal lipid metabolism and the relationship between tumor invasion and metastasis. Front Med. 2021;15(5):649–56.

    Article  PubMed  Google Scholar 

  140. Gonzalez-Salinas F, Rojo R, Martinez-Amador C, Herrera-Gamboa J, Trevino V. Transcriptomic and cellular analyses of CRISPR/Cas9-mediated edition of FASN show inhibition of aggressive characteristics in breast cancer cells. Biochem Biophys Res Commun. 2020;529(2):321–7.

    Article  CAS  PubMed  Google Scholar 

  141. Lenoir WF, Morgado M, DeWeirdt PC, McLaughlin M, Griffith AL, Sangree AK, Feeley MN, Esmaeili Anvar N, Kim E, Bertolet LL. Discovery of putative tumor suppressors from CRISPR screens reveals rewired lipid metabolism in acute myeloid leukemia cells. Nat Commun. 2021;12(1):1–15.

    Article  CAS  Google Scholar 

  142. Saha B, Mathur T, Tronolone JJ, Chokshi M, Lokhande GK, Selahi A, Gaharwar AK, Afshar-Kharghan V, Sood AK, Bao G. Human tumor microenvironment chip evaluates the consequences of platelet extravasation and combinatorial antitumor-antiplatelet therapy in ovarian cancer. Sci Adv. 2021;7(30):eabg5283.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Asghar S, Parvaiz F, Manzoor S. Multifaceted role of cancer educated platelets in survival of cancer cells. Thromb Res. 2019;177:42–50.

    Article  CAS  PubMed  Google Scholar 

  144. Mammadova-Bach E, Gil-Pulido J, Sarukhanyan E, Burkard P, Shityakov S, Schonhart C, Stegner D, Remer K, Nurden P, Nurden AT. Platelet glycoprotein VI promotes metastasis through interaction with cancer cell–derived galectin-3. Blood. 2020;135(14):1146–60.

    PubMed  Google Scholar 

  145. Volz J, Mammadova-Bach E, Gil-Pulido J, Nandigama R, Remer K, Sorokin L, Zernecke A, Abrams SI, Ergün S, Henke E. Inhibition of platelet GPVI induces intratumor hemorrhage and increases efficacy of chemotherapy in mice. Blood J Am Soc Hematol. 2019;133(25):2696–706.

    CAS  Google Scholar 

  146. Tu CF, Wu MY, Lin YC, Kannagi R, Yang RB. FUT8 promotes breast cancer cell invasiveness by remodeling TGF-β receptor core fucosylation. Breast Cancer Res. 2017;19(1):111.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  147. Wang Y, Li Y, Ma Y, Wu W. DCLK1 promotes malignant progression of breast cancer by regulating Wnt/β-catenin signaling pathway. Eur Rev Med Pharmacol Sci. 2019;23(21):9489–98.

    PubMed  Google Scholar 

  148. Liu H, Wen T, Zhou Y, Fan X, Du T, Gao T, Li L, Liu J, Yang L, Yao J. DCLK1 plays a metastatic-promoting role in human breast cancer cells. BioMed Res Int. 2019;2019:1.

    Google Scholar 

  149. Overgaard J, Eriksen JG, Nordsmark M, Alsner J, Horsman MR. Plasma osteopontin, hypoxia, and response to the hypoxia sensitiser nimorazole in radiotherapy of head and neck cancer: results from the DAHANCA 5 randomised double-blind placebo-controlled trial. Lancet Oncol. 2005;6(10):757–64.

    Article  CAS  PubMed  Google Scholar 

  150. Wang M, Han J, Marcar L, Black J, Liu Q, Li X, Nagulapalli K, Sequist LV, Mak RH, Benes CH. Radiation resistance in KRAS-mutated lung cancer is enabled by stem-like properties mediated by an osteopontin–EGFR pathway. Cancer Res. 2017;77(8):2018–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Behbahani R, Danyaei A, Teimoori A, Neisi N, Tahmasbi M. Breast cancer radioresistance may be overcome by osteopontin gene knocking out with CRISPR/Cas9 technique. Cancer/Radiothérapie. 2021;25(3):222–8.

    Article  CAS  Google Scholar 

  152. Ehrke-Schulz E, Heinemann S, Schulte L, Schiwon M, Ehrhardt A. Adenoviral vectors armed with PAPILLOMAVIRUs oncogene specific CRISPR/Cas9 kill human-papillomavirus-induced cervical cancer cells. Cancers. 2020;12(7):1934.

    Article  CAS  PubMed Central  Google Scholar 

  153. Park JS, Kim EJ, Lee JY, Sin HS, Namkoong SE, Um SJ. Functional inactivation of p73, a homolog of p53 tumor suppressor protein, by human papillomavirus E6 proteins. Int J Cancer. 2001;91(6):822–7.

    Article  CAS  PubMed  Google Scholar 

  154. Ding W, Hu Z, Zhu D, Jiang X, Yu L, Wang X, Zhang C, Wang L, Ji T, Li K. Zinc finger nucleases targeting the human papillomavirus E7 oncogene induce E7 disruption and a transformed phenotype in HPV16/18-positive cervical cancer cells. Clin Cancer Res. 2014;20(24):6495–503.

    Article  CAS  PubMed  Google Scholar 

  155. Shankar S, Prasad D, Sanawar R, Das AV, Pillai MR. TALEN based HPV-E7 editing triggers necrotic cell death in cervical cancer cells. Sci Rep. 2017;7(1):1–12.

    Article  CAS  Google Scholar 

  156. Herfs M, Herman L, Hubert P, Minner F, Arafa M, Roncarati P, Henrotin Y, Boniver J, Delvenne P. High expression of PGE2 enzymatic pathways in cervical (pre) neoplastic lesions and functional consequences for antigen-presenting cells. Cancer Immunol Immunother. 2009;58(4):603–14.

    Article  CAS  PubMed  Google Scholar 

  157. Hojnik M, Frković Grazio S, Verdenik I, Rižner TL. AKR1B1 and AKR1B10 as prognostic biomarkers of endometrioid endometrial carcinomas. Cancers. 2021;13(14):3398.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Lacroix Pépin N, Chapdelaine P, Rodriguez Y, Tremblay J-P, Fortier MA. Generation of human endometrial knockout cell lines with the CRISPR/Cas9 system confirms the prostaglandin F2α synthase activity of aldo-ketoreductase 1B1. Mol Hum Reprod. 2014;20(7):650–63.

    Article  PubMed  CAS  Google Scholar 

  159. Ji J, Xu M-X, Qian T-Y, Zhu S-Z, Jiang F, Liu Z-X, Xu W-S, Zhou J, Xiao M-B. The AKR1B1 inhibitor epalrestat suppresses the progression of cervical cancer. Mol Biol Rep. 2020;47(8):6091–103.

    Article  CAS  PubMed  Google Scholar 

  160. Huang J, Diao G, Zhang Q, Chen Y, Han J, Guo J. E6-regulated overproduction of prostaglandin E2 may inhibit migration of dendritic cells in human papillomavirus 16-positive cervical lesions. Int J Oncol. 2020;56(4):921–31.

    CAS  PubMed  PubMed Central  Google Scholar 

  161. Mo X-T, Leung TH-Y, Tang HW-M, Siu MK-Y, Wan PK-T, Chan KK-L, Cheung AN-Y, Ngan HY-S. CD109 mediates tumorigenicity and cancer aggressiveness via regulation of EGFR and STAT3 signalling in cervical squamous cell carcinoma. Br J Cancer. 2020;123(5):833–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Zhang JM, Hashimoto M, Kawai K, Murakumo Y, Sato T, Ichihara M, Nakamura S, Takahashi M. CD109 expression in squamous cell carcinoma of the uterine cervix. Pathol Int. 2005;55(4):165–9.

    Article  CAS  PubMed  Google Scholar 

  163. Zeinalzadeh E, Valerievich Yumashev A, Rahman HS, Marofi F, Shomali N, Kafil HS, Solali S, Sajjadi-Dokht M, Vakili-Samiani S, Jarahian M, et al. The role of janus kinase/STAT3 pathway in hematologic malignancies with an emphasis on epigenetics. Front Genet. 2021;12:703883.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Han Y-Q, Ming S-L, Wu H-T, Zeng L, Ba G, Li J, Lu W-F, Han J, Du Q-J, Sun M-M. Myostatin knockout induces apoptosis in human cervical cancer cells via elevated reactive oxygen species generation. Redox Biol. 2018;19:412–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Curran KJ, Pegram HJ, Brentjens RJ. Chimeric antigen receptors for T cell immunotherapy: current understanding and future directions. J Gene Med. 2012;14(6):405–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Shen Y, Chen F, Liang Y. MicroRNA-133a inhibits the proliferation of non-small cell lung cancer by targeting YES1. Oncol Lett. 2019;18(6):6759–65.

    CAS  PubMed  PubMed Central  Google Scholar 

  167. Zhu H, Lu Q, Lu Q, Shen X, Yu L. Matrine regulates proliferation, apoptosis, cell cycle, migration, and invasion of non-small cell lung cancer cells through the circFUT8/miR-944/YES1 axis. Cancer Manag Res. 2021;13:3429.

    Article  PubMed  PubMed Central  Google Scholar 

  168. Garmendia I, Pajares MJ, Hermida-Prado F, Ajona D, Bértolo C, Sainz C, Lavín A, Remírez AB, Valencia K, Moreno H. YES1 drives lung cancer growth and progression and predicts sensitivity to dasatinib. Am J Respir Crit Care Med. 2019;200(7):888–99.

    Article  CAS  PubMed  Google Scholar 

  169. Bilal E, Alexe G, Yao M, Cong L, Kulkarni A, Ginjala V, Toppmeyer D, Ganesan S, Bhanot G. Identification of the YES1 kinase as a therapeutic target in basal-like breast cancers. Genes Cancer. 2010;1(10):1063–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Zhou Y, Wang C, Ding J, Chen Y, Sun Y, Cheng Z. miR-133a targets YES1 to reduce cisplatin resistance in ovarian cancer by regulating cell autophagy. Cancer Cell Int. 2022;22(1):1–13.

    Article  CAS  Google Scholar 

  171. Zhang L, Yang Y, Chai L, Bu H, Yang Y, Huang H, Ran J, Zhu Y, Li L, Chen F. FRK plays an oncogenic role in non-small cell lung cancer by enhancing the stemness phenotype via induction of metabolic reprogramming. Int J Cancer. 2020;146(1):208–22.

    Article  CAS  PubMed  Google Scholar 

  172. Zang Q, Xu L, Li J, Jia H. GATA6 activated long non-coding RNA PCAT1 maintains stemness of non-small cell lung cancer by mediating FRK. J BUON. 2020;25(5):2371–81.

    PubMed  Google Scholar 

  173. Li L, Kou Y, Chen F, Sun X. Expression of p-FRK and its prognostic analysis in non-small cell lung cancer patients. Chin J Clin Exp Pathol. 2017:525–9.

  174. Grunblatt E, Wu N, Zhang H, Liu X, Norton JP, Ohol Y, Leger P, Hiatt JB, Eastwood EC, Thomas R. MYCN drives chemoresistance in small cell lung cancer while USP7 inhibition can restore chemosensitivity. Genes Dev. 2020;34(17–18):1210–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Nau MM, Brooks BJ, Carney DN, Gazdar AF, Battey JF, Sausville EA, Minna JD. Human small-cell lung cancers show amplification and expression of the N-myc gene. Proc Natl Acad Sci. 1986;83(4):1092–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Czech-Sioli M, Siebels S, Radau S, Zahedi RP, Schmidt C, Dobner T, Grundhoff A, Fischer N. The ubiquitin-specific protease Usp7, a novel Merkel cell polyomavirus large T-antigen interaction partner, modulates viral DNA replication. J Virol. 2020;94(5):e01638-e1719.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Bhattacharya S, Chakraborty D, Basu M, Ghosh MK. Emerging insights into HAUSP (USP7) in physiology, cancer and other diseases. Signal Transduct Target Ther. 2018;3(1):1–12.

    Article  CAS  Google Scholar 

  178. Weinstock J, Wu J, Cao P, Kingsbury WD, McDermott JL, Kodrasov MP, McKelvey DM, Suresh Kumar K, Goldenberg SJ, Mattern MR. Selective dual inhibitors of the cancer-related deubiquitylating proteases USP7 and USP47. ACS Med Chem Lett. 2012;3(10):789–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Eser S, Schnieke A, Schneider G, Saur D. Oncogenic KRAS signalling in pancreatic cancer. Br J Cancer. 2014;111(5):817–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Lentsch E, Li L, Pfeffer S, Ekici AB, Taher L, Pilarsky C, Grützmann R. CRISPR/Cas9-mediated knock-out of krasG12D mutated pancreatic cancer cell lines. Int J Mol Sci. 2019;20(22):5706.

    Article  CAS  PubMed Central  Google Scholar 

  181. Ischenko I, D’Amico S, Rao M, Li J, Hayman MJ, Powers S, Petrenko O, Reich NC. KRAS drives immune evasion in a genetic model of pancreatic cancer. Nat Commun. 2021;12(1):1–15.

    Article  CAS  Google Scholar 

  182. Tao J, Yang G, Zhou W, Qiu J, Chen G, Luo W, Zhao F, You L, Zheng L, Zhang T. Targeting hypoxic tumor microenvironment in pancreatic cancer. J Hematol Oncol. 2021;14(1):1–25.

    Article  Google Scholar 

  183. Wei H, Li F, Fu P, Liu X. Effects of the silencing of hypoxia-inducible factor-1 alpha on metastasis of pancreatic cancer. Eur Rev Med Pharmacol Sci. 2013;17(4):436–46.

    CAS  PubMed  Google Scholar 

  184. Li M, Xie H, Liu Y, Xia C, Cun X, Long Y, Chen X, Deng M, Guo R, Zhang Z. Knockdown of hypoxia-inducible factor-1 alpha by tumor targeted delivery of CRISPR/Cas9 system suppressed the metastasis of pancreatic cancer. J Control Release. 2019;304:204–15.

    Article  CAS  PubMed  Google Scholar 

  185. Wei X, Yang J, Adair SJ, Ozturk H, Kuscu C, Lee KY, Kane WJ, O’Hara PE, Liu D, Demirlenk YM. Targeted CRISPR screening identifies PRMT5 as synthetic lethality combinatorial target with gemcitabine in pancreatic cancer cells. Proc Natl Acad Sci. 2020;117(45):28068–79.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Yasunaga K, Ito T, Miki M, Ueda K, Fujiyama T, Tachibana Y, Fujimori N, Kawabe K, Ogawa Y. Using CRISPR/Cas9 to knock out amylase in acinar cells decreases pancreatitis-induced autophagy. BioMed Res Int. 2018;2018:1.

    Article  CAS  Google Scholar 

  187. Hwang M, Jun DW, Kang EH, Yoon K-A, Cheong H, Kim Y-H, Lee C-H, Kim S. EI24, as a component of autophagy, is involved in pancreatic cell proliferation. Front Oncol. 2019;9:652.

    Article  PubMed  PubMed Central  Google Scholar 

  188. Zang Y, Zhu L, Li T, Wang Q, Li J, Qian Y, Wei L, Xie M, Tang W-H, Liu X. EI24 Suppresses tumorigenesis in pancreatic cancer via regulating c-Myc. Gastroenterol Res Pract. 2018;2018:1.

    Article  Google Scholar 

  189. Uzzo RG, Crispen PL, Golovine K, Makhov P, Horwitz EM, Kolenko VM. Diverse effects of zinc on NF-κB and AP-1 transcription factors: implications for prostate cancer progression. Carcinogenesis. 2006;27(10):1980–90.

    Article  CAS  PubMed  Google Scholar 

  190. Sato N, Sadar MD, Bruchovsky N, Saatcioglu F, Rennie PS, Sato S, Lange PH, Gleave ME. Androgenic induction of prostate-specific antigen gene is repressed by protein-protein interaction between the androgen receptor and AP-1/c-Jun in the human prostate cancer cell line LNCaP. J Biol Chem. 1997;272(28):17485–94.

    Article  CAS  PubMed  Google Scholar 

  191. Ouyang X, Jessen WJ, Al-Ahmadie H, Serio AM, Lin Y, Shih W-J, Reuter VE, Scardino PT, Shen MM, Aronow BJ. Activator protein-1 transcription factors are associated with progression and recurrence of prostate cancer. Cancer Res. 2008;68(7):2132–44.

    Article  CAS  PubMed  Google Scholar 

  192. Riedel M, Berthelsen MF, Cai H, Haldrup J, Borre M, Paludan SR, Hager H, Vendelbo MH, Wagner EF, Bakiri L. In vivo CRISPR inactivation of Fos promotes prostate cancer progression by altering the associated AP-1 subunit Jun. Oncogene. 2021;40(13):2437–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Riedel M, Cai H, Stoltze IC, Vendelbo MH, Wagner EF, Bakiri L, Thomsen MK. Targeting AP-1 transcription factors by CRISPR in the prostate. Oncotarget. 2021;12(19):1956.

    Article  PubMed  PubMed Central  Google Scholar 

  194. Pi M, Quarles LD. GPRC6A regulates prostate cancer progression. Prostate. 2012;72(4):399–409.

    Article  CAS  PubMed  Google Scholar 

  195. Ye R, Pi M, Cox JV, Nishimoto SK, Quarles LD. CRISPR/Cas9 targeting of GPRC6A suppresses prostate cancer tumorigenesis in a human xenograft model. J Exp Clin Cancer Res. 2017;36(1):1–13.

    Article  CAS  Google Scholar 

  196. Chakraborty G, Patail NK, Hirani R, Nandakumar S, Mazzu YZ, Yoshikawa Y, Atiq M, Jehane LE, Stopsack KH, Lee G-SM. Attenuation of SRC kinase activity augments PARP inhibitor–mediated synthetic lethality in BRCA2-altered prostate tumors. Clin Cancer Res. 2021;27(6):1792–806.

    Article  CAS  PubMed  Google Scholar 

  197. Su B, Zhang L, Zhuang W, Zhang W, Chen X. Knockout of Akt1/2 suppresses the metastasis of human prostate cancer cells CWR22rv1 in vitro and in vivo. J Cell Mol Med. 2021;25(3):1546–53.

    Article  CAS  PubMed  Google Scholar 

  198. Huang XF, Chen JZ. Obesity, the PI3K/Akt signal pathway and colon cancer. Obes Rev. 2009;10(6):610–6.

    Article  CAS  PubMed  Google Scholar 

  199. Song H, Xu Y, Xu T, Fan R, Jiang T, Cao M, Shi L, Song J. CircPIP5K1A activates KRT80 and PI3K/AKT pathway to promote gastric cancer development through sponging miR-671–5p. Biomed Pharmacother. 2020;126:109941.

    Article  CAS  PubMed  Google Scholar 

  200. Stemke-Hale K, Gonzalez-Angulo AM, Lluch A, Neve RM, Kuo W-L, Davies M, Carey M, Hu Z, Guan Y, Sahin A. An integrative genomic and proteomic analysis of PIK3CA, PTEN, and AKT mutations in breast cancer. Cancer Res. 2008;68(15):6084–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Pérez-Tenorio G, Stål O. Activation of AKT/PKB in breast cancer predicts a worse outcome among endocrine treated patients. Br J Cancer. 2002;86(4):540–5.

    Article  PubMed  PubMed Central  Google Scholar 

  202. Fang W, Huang Y, Gu W, Gan J, Wang W, Zhang S, Wang K, Zhan J, Yang Y, Huang Y. PI3K-AKT-mTOR pathway alterations in advanced NSCLC patients after progression on EGFR-TKI and clinical response to EGFR-TKI plus everolimus combination therapy. Transl Lung Cancer Res. 2020;9(4):1258.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Ediriweera MK, Tennekoon KH, Samarakoon SR, editors. Role of the PI3K/AKT/mTOR signaling pathway in ovarian cancer: Biological and therapeutic significance. Seminars in cancer biology; 2019: Elsevier.

  204. Rahmani F, Ziaeemehr A, Shahidsales S, Gharib M, Khazaei M, Ferns GA, Ryzhikov M, Avan A, Hassanian SM. Role of regulatory miRNAs of the PI3K/AKT/mTOR signaling in the pathogenesis of hepatocellular carcinoma. J Cell Physiol. 2020;235(5):4146–52.

    Article  CAS  PubMed  Google Scholar 

  205. Kalli M, Minia A, Pliaka V, Fotis C, Alexopoulos LG, Stylianopoulos T. Solid stress-induced migration is mediated by GDF15 through Akt pathway activation in pancreatic cancer cells. Sci Rep. 2019;9(1):1–12.

    Article  CAS  Google Scholar 

  206. Nasrollahzadeh A, Momeny M, Fasehee H, Yaghmaie M, Bashash D, Hassani S, Mousavi SA, Ghaffari SH. Anti-proliferative activity of disulfiram through regulation of the AKT-FOXO axis: a proteomic study of molecular targets. Biochim Biophys Acta Mol Cell Res 2021;1868(10):119087.

    Article  CAS  PubMed  Google Scholar 

  207. Zhu C-c, Chen C, Xu Z-q, Zhao J-k, Ou B-c, Sun J, Zheng M-h, Zong Y-p, Lu A-g. CCR6 promotes tumor angiogenesis via the AKT/NF-κB/VEGF pathway in colorectal cancer. Biochim Biophys Acta Mol Basis Dis 2018;1864(2):387–97.

    Article  CAS  PubMed  Google Scholar 

  208. Nabipoorashrafi SA, Shomali N, Sadat-Hatamnezhad L, Mahami-Oskouei M, Mahmoudi J, Sandoghchian Shotorbani B, Akbari M, Xu H, Sandoghchian SS. miR-143 acts as an inhibitor of migration and proliferation as well as an inducer of apoptosis in melanoma cancer cells in vitro. IUBMB Life. 2020;72(9):2034–44.

    Article  CAS  PubMed  Google Scholar 

  209. Lin K, Shen S-H, Lu F, Zheng P, Wu S, Liao J, Jiang X, Zeng G, Wei D. CRISPR screening of E3 ubiquitin ligases reveals Ring Finger Protein 185 as a novel tumor suppressor in glioblastoma repressed by promoter hypermethylation and miR-587. J Transl Med. 2022;20(1):1–14.

    Article  CAS  Google Scholar 

  210. Gruffaz M, Yuan H, Meng W, Liu H, Bae S, Kim J-S, Lu C, Huang Y, Gao S-J. CRISPR-Cas9 screening of Kaposi’s sarcoma-associated herpesvirus-transformed cells identifies XPO1 as a vulnerable target of cancer cells. MBio. 2019;10(3):e00866-e919.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Zhang Y-Q, Pei J-H, Shi S-S, Guo X-s, Cui G-y, Li Y-F, Zhang H-P, Hu W-Q. CRISPR/Cas9-mediated knockout of the PDEF gene inhibits migration and invasion of human gastric cancer AGS cells. Biomed Pharmacother. 2019;111:76–85.

    Article  CAS  PubMed  Google Scholar 

  212. Koepsell H. The Na+-D-glucose cotransporters SGLT1 and SGLT2 are targets for the treatment of diabetes and cancer. Pharmacol Ther. 2017;170:148–65.

    Article  CAS  PubMed  Google Scholar 

  213. Liu H, Ertay A, Peng P, Li J, Liu D, Xiong H, Zou Y, Qiu H, Hancock D, Yuan X. SGLT1 is required for the survival of triple-negative breast cancer cells via potentiation of EGFR activity. Mol Oncol. 2019;13(9):1874–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Shi M, Wang C, Ji J, Cai Q, Zhao Q, Xi W, Zhang J. CRISPR/Cas9-mediated knockout of SGLT1 inhibits proliferation and alters metabolism of gastric cancer cells. Cell Signalling. 2022;90:110192.

    Article  CAS  PubMed  Google Scholar 

  215. Haghighi N, Doosti A, Kiani J. Evaluation of CRISPR/Cas9 system effects on knocking out NEAT1 gene in AGS gastric cancer cell line with therapeutic perspective. J Gastrointestinal Cancer. 2021:1–9.

  216. Yu X, Li Z, Zheng H, Chan MT, Wu WKK. NEAT 1: a novel cancer-related long non-coding RNA. Cell Prolif. 2017;50(2):e12329.

    Article  PubMed Central  CAS  Google Scholar 

  217. Chen Z, Zhang Z, Xie B, Zhang H. Clinical significance of up-regulated lncRNA NEAT1 in prognosis of ovarian cancer. Eur Rev Med Pharmacol Sci. 2016;20(16):3373–7.

    PubMed  Google Scholar 

  218. Choi BD, Archer GE, Mitchell DA, Heimberger AB, McLendon RE, Bigner DD, Sampson JH. EGFRvIII-targeted vaccination therapy of malignant glioma. Brain Pathol. 2009;19(4):713–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Morgan RA, Johnson LA, Davis JL, Zheng Z, Woolard KD, Reap EA, Feldman SA, Chinnasamy N, Kuan C-T, Song H. Recognition of glioma stem cells by genetically modified T cells targeting EGFRvIII and development of adoptive cell therapy for glioma. Hum Gene Ther. 2012;23(10):1043–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  220. Huang K, Yang C, Wang Q-x, Li Y-s, Fang C, Tan Y-l, Wei J-w, Wang Y-f, Li X, Zhou J-h. The CRISPR/Cas9 system targeting EGFR exon 17 abrogates NF-κB activation via epigenetic modulation of UBXN1 in EGFRwt/vIII glioma cells. Cancer Lett. 2017;388:269–80.

    Article  CAS  PubMed  Google Scholar 

  221. Ding H, Inoue S, Ljubimov AV, Patil R, Portilla-Arias J, Konda B, Black KL, Holler E, Ljubimova JY. Inhibition of tumor vascular protein laminin-411 by nanobioconjugate for glioma treatment. AACR; 2011.

  222. Klymyshyn D, Galstyan A, Patil R, Ding H, Shatalova E, Wagner S, Black K, Ljubimov A, Holler E, Ljubimova J. Blockade of laminin-411-notch crosstalk as an effective therapy for glioblastoma treatment. AACR; 2019.

  223. Ljubimova J, Inoue S, Bannykh S, Phuphanich S, Rudnick J, Ljubimov A, Black K. Prognostic significance of new glioma marker, laminin-411, and its inhibition by targeted nanobiopolymer in vivo. J Clin Oncol. 2011;29(15_suppl):e13008.

    Article  Google Scholar 

  224. Sun T, Patil R, Galstyan A, Klymyshyn D, Ding H, Chesnokova A, Cavenee WK, Furnari FB, Ljubimov VA, Shatalova ES. Blockade of a laminin-411–notch axis with CRISPR/Cas9 or a nanobioconjugate inhibits glioblastoma growth through tumor-microenvironment cross-talk. Cancer Res. 2019;79(6):1239–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Hong M, Clubb JD, Chen YY. Engineering CAR-T cells for next-generation cancer therapy. Cancer Cell. 2020;38(4):473–88.

    Article  CAS  PubMed  Google Scholar 

  226. Glienke W, Esser R, Priesner C, Suerth JD, Schambach A, Wels WS, Grez M, Kloess S, Arseniev L, Koehl U. Advantages and applications of CAR-expressing natural killer cells. Front Pharmacol. 2015;6:21.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  227. Haji-Fatahaliha M, Hosseini M, Akbarian A, Sadreddini S, Jadidi-Niaragh F, Yousefi M. CAR-modified T-cell therapy for cancer: an updated review. Artif Cells Nanomed Biotechnol. 2016;44(6):1339–49.

    Article  CAS  PubMed  Google Scholar 

  228. Han D, Xu Z, Zhuang Y, Ye Z, Qian Q. Current progress in CAR-T cell therapy for hematological malignancies. J Cancer. 2021;12(2):326–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  229. Qu J, Mei Q, Chen L, Zhou J. Chimeric antigen receptor (CAR)-T-cell therapy in non-small-cell lung cancer (NSCLC): current status and future perspectives. Cancer Immunol Immunother. 2021;70(3):619–31.

    Article  CAS  PubMed  Google Scholar 

  230. Corti C, Venetis K, Sajjadi E, Zattoni L, Curigliano G, Fusco N. CAR-T cell therapy for triple-negative breast cancer and other solid tumors: preclinical and clinical progress. Expert Opin Investig Drugs. 2022:1–13.

  231. Razeghian E, Nasution MK, Rahman HS, Gardanova ZR, Abdelbasset WK, Aravindhan S, Bokov DO, Suksatan W, Nakhaei P, Shariatzadeh S. A deep insight into CRISPR/Cas9 application in CAR-T cell-based tumor immunotherapies. Stem Cell Res Ther. 2021;12(1):1–17.

    Article  CAS  Google Scholar 

  232. Zhao J, Lin Q, Song Y, Liu D. Universal CARs, universal T cells, and universal CAR T cells. J Hematol Oncol. 2018;11(1):1–9.

    Article  CAS  Google Scholar 

  233. Budi HS, Ahmad FN, Achmad H, Ansari MJ, Mikhailova MV, Suksatan W, Chupradit S, Shomali N, Marofi F. Human epidermal growth factor receptor 2 (HER2)-specific chimeric antigen receptor (CAR) for tumor immunotherapy; recent progress. Stem Cell Res Ther. 2022;13(1):1–21.

    Article  CAS  Google Scholar 

  234. Marofi F, Tahmasebi S, Rahman HS, Kaigorodov D, Markov A, Yumashev AV, Shomali N, Chartrand MS, Pathak Y, Mohammed RN, et al. Correction to: any closer to successful therapy of multiple myeloma? CAR-T cell is a good reason for optimism. Stem Cell Res Ther. 2021;12(1):443.

    Article  PubMed  PubMed Central  Google Scholar 

  235. Cherkassky L, Morello A, Villena-Vargas J, Feng Y, Dimitrov DS, Jones DR, Sadelain M, Adusumilli PS. Human CAR T cells with cell-intrinsic PD-1 checkpoint blockade resist tumor-mediated inhibition. J Clin Investig. 2016;126(8):3130–44.

    Article  PubMed  PubMed Central  Google Scholar 

  236. Marofi F, Rahman HS, Thangavelu L, Dorofeev A, Bayas-Morejón F, Shirafkan N, Shomali N, Chartrand MS, Jarahian M, Vahedi G. Renaissance of armored immune effector cells, CAR-NK cells, brings the higher hope for successful cancer therapy. Stem Cell Res Ther. 2021;12(1):1–21.

    Article  CAS  Google Scholar 

  237. June CH, O’Connor RS, Kawalekar OU, Ghassemi S, Milone MC. CAR T cell immunotherapy for human cancer. Science. 2018;359(6382):1361–5.

    Article  CAS  PubMed  Google Scholar 

  238. Liu X, Zhang Y, Cheng C, Cheng AW, Zhang X, Li N, Xia C, Wei X, Liu X, Wang H. CRISPR-Cas9-mediated multiplex gene editing in CAR-T cells. Cell Res. 2017;27(1):154–7.

    Article  PubMed  CAS  Google Scholar 

  239. McCreedy BJ, Senyukov VV, Nguyen KT. Off the shelf T cell therapies for hematologic malignancies. Best Pract Res Clin Haematol. 2018;31(2):166–75.

    Article  PubMed  Google Scholar 

  240. Cooper ML, Choi J, Staser K, Ritchey JK, Devenport JM, Eckardt K, Rettig MP, Wang B, Eissenberg LG, Ghobadi A. An, “off-the-shelf” fratricide-resistant CAR-T for the treatment of T cell hematologic malignancies. Leukemia. 2018;32(9):1970–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  241. Kagoya Y, Guo T, Yeung B, Saso K, Anczurowski M, Wang C-H, Murata K, Sugata K, Saijo H, Matsunaga Y. Genetic ablation of HLA class I, class II, and the T-cell receptor enables allogeneic T cells to be used for adoptive T-cell therapy. Cancer Immunol Res. 2020;8(7):926–36.

    Article  CAS  PubMed  Google Scholar 

  242. Hu B, Zou Y, Zhang L, Tang J, Niedermann G, Firat E, Huang X, Zhu X. Nucleofection with plasmid DNA for CRISPR/Cas9-mediated inactivation of programmed cell death protein 1 in CD133-specific CAR T cells. Hum Gene Ther. 2019;30(4):446–58.

    Article  CAS  PubMed  Google Scholar 

  243. Nakazawa T, Natsume A, Nishimura F, Morimoto T, Matsuda R, Nakamura M, Yamada S, Nakagawa I, Motoyama Y, Park Y-S, et al. Effect of CRISPR/Cas9-mediated PD-1-disrupted primary human third-generation CAR-T cells targeting EGFRvIII on in vitro human glioblastoma cell growth. Cells. 2020;9(4):998.

    Article  CAS  PubMed Central  Google Scholar 

  244. Zhang Y, Zhang X, Cheng C, Mu W, Liu X, Li N, Wei X, Liu X, Xia C, Wang H. CRISPR-Cas9 mediated LAG-3 disruption in CAR-T cells. Front Med. 2017;11(4):554–62.

    Article  PubMed  Google Scholar 

  245. Tang N, Cheng C, Zhang X, Qiao M, Li N, Mu W, Wei X-F, Han W, Wang H. TGF-β inhibition via CRISPR promotes the long-term efficacy of CAR T cells against solid tumors. JCI insight 2020;5(4):e133977.

    Article  PubMed Central  Google Scholar 

  246. Welstead GG, Vong Q, Nye C, Hause R, Clouser C, Jones J, Burleigh S, Borges CM, Chin M, Marco E, editors. Improving efficacy of CAR T cells through CRISPR/Cas9 mediated knockout of TGFBR2. Molecular Therapy; 2018: CELL PRESS 50 HAMPSHIRE ST, FLOOR 5, CAMBRIDGE, MA 02139 USA.

  247. Kloss CC, Lee J, Zhang A, Chen F, Melenhorst JJ, Lacey SF, Maus MV, Fraietta JA, Zhao Y, June CH. Dominant-negative TGF-β receptor enhances PSMA-targeted human CAR T cell proliferation and augments prostate cancer eradication. Mol Ther. 2018;26(7):1855–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Sterner RM, Sakemura R, Cox MJ, Yang N, Khadka RH, Forsman CL, Hansen MJ, Jin F, Ayasoufi K, Hefazi M. GM-CSF inhibition reduces cytokine release syndrome and neuroinflammation but enhances CAR-T cell function in xenografts. Blood J Am Soc Hematol. 2019;133(7):697–709.

    CAS  Google Scholar 

  249. Khadka RH, Sakemura R, Kenderian SS, Johnson AJ. Management of cytokine release syndrome: an update on emerging antigen-specific T cell engaging immunotherapies. Immunotherapy. 2019;11(10):851–7.

    Article  CAS  PubMed  Google Scholar 

  250. Sterner RM, Cox MJ, Sakemura R, Kenderian SS. Using CRISPR/Cas9 to knock out GM-CSF in CAR-T cells. J Vis Exp. 2019(149).

  251. Yi Y, Chai X, Zheng L, Zhang Y, Shen J, Hu B, Tao G. CRISPR-edited CART with GM-CSF knockout and auto secretion of IL6 and IL1 blockers in patients with hematologic malignancy. Cell Discovery. 2021;7(1):1–11.

    Article  CAS  Google Scholar 

  252. van Dongen JE, Berendsen JT, Steenbergen RD, Wolthuis RM, Eijkel JC, Segerink LI. Point-of-care CRISPR/Cas nucleic acid detection: recent advances, challenges and opportunities. Biosensors Bioelectronics. 2020;166:112445.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  253. Cho SW, Kim S, Kim Y, Kweon J, Kim HS, Bae S, Kim J-S. Analysis of off-target effects of CRISPR/Cas-derived RNA-guided endonucleases and nickases. Genome Res. 2014;24(1):132–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  254. Baghini SS, Gardanova ZR, Zekiy AO, Shomali N, Tosan F, Jarahian M. Optimizing sgRNA to improve CRISPR/Cas9 knockout efficiency: special focus on human and animal cell. Front Bioeng Biotechnol. 2021;9.

  255. Manghwar H, Li B, Ding X, Hussain A, Lindsey K, Zhang X, Jin S. CRISPR/Cas systems in genome editing: methodologies and tools for sgRNA design, off-target evaluation, and strategies to mitigate off-target effects. Adv Sci. 2020;7(6):1902312.

    Article  CAS  Google Scholar 

  256. Wolt JD, Wang K, Sashital D, Lawrence‐Dill CJ. Achieving plant CRISPR targeting that limits off‐target effects. Plant Genome. 2016;9(3):plantgenome2016.05.0047.

  257. Liu X, Homma A, Sayadi J, Yang S, Ohashi J, Takumi T. Sequence features associated with the cleavage efficiency of CRISPR/Cas9 system. Sci Rep. 2016;6(1):1–9.

    CAS  Google Scholar 

  258. Zhang X-H, Tee LY, Wang X-G, Huang Q-S, Yang S-H. Off-target effects in CRISPR/Cas9-mediated genome engineering. Mol Ther Nucleic Acids 2015;4:e264.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  259. Guilinger JP, Thompson DB, Liu DR. Fusion of catalytically inactive Cas9 to FokI nuclease improves the specificity of genome modification. Nat Biotechnol. 2014;32(6):577–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  260. Ran FA, Hsu PD, Lin C-Y, Gootenberg JS, Konermann S, Trevino AE, Scott DA, Inoue A, Matoba S, Zhang Y. Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity. Cell. 2013;154(6):1380–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  261. Tsai SQ, Wyvekens N, Khayter C, Foden JA, Thapar V, Reyon D, Goodwin MJ, Aryee MJ, Joung JK. Dimeric CRISPR RNA-guided FokI nucleases for highly specific genome editing. Nat Biotechnol. 2014;32(6):569–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  262. Singh R, Kuscu C, Quinlan A, Qi Y, Adli M. Cas9-chromatin binding information enables more accurate CRISPR off-target prediction. Nucleic Acids Res. 2015;43(18):e118.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  263. Tsai SQ, Nguyen NT, Malagon-Lopez J, Topkar VV, Aryee MJ, Joung JK. CIRCLE-seq: a highly sensitive in vitro screen for genome-wide CRISPR-Cas9 nuclease off-targets. Nat Methods. 2017;14(6):607–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  264. Barrangou R, Fremaux C, Deveau H, Richards M, Boyaval P, Moineau S, Romero DA, Horvath P. CRISPR provides acquired resistance against viruses in prokaryotes. Science. 2007;315(5819):1709–12.

    Article  CAS  PubMed  Google Scholar 

  265. Cong L, Ran FA, Cox D, Lin S, Barretto R, Habib N, Hsu PD, Wu X, Jiang W, Marraffini LA, et al. Multiplex genome engineering using CRISPR/Cas systems. Science. 2013;339(6121):819–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  266. Slipek NJ, Varshney J, Largaespada DA. CRISPR/Cas9-based positive screens for cancer-related traits. Methods Mol Biol. 2019;1907:137–44.

    Article  CAS  PubMed  Google Scholar 

  267. He C, Han S, Chang Y, Wu M, Zhao Y, Chen C, Chu X. CRISPR screen in cancer: status quo and future perspectives. Am J Cancer Res. 2021;11(4):1031–50.

    CAS  PubMed  PubMed Central  Google Scholar 

  268. Adelmann CH, Wang T, Sabatini DM, Lander ES. Genome-wide CRISPR/Cas9 screening for identification of cancer genes in cell lines. Methods Mol Biol. 2019;1907:125–36.

    Article  CAS  PubMed  Google Scholar 

  269. Kleinstiver BP, Prew MS, Tsai SQ, Topkar VV, Nguyen NT, Zheng Z, Gonzales AP, Li Z, Peterson RT, Yeh JR, et al. Engineered CRISPR-Cas9 nucleases with altered PAM specificities. Nature. 2015;523(7561):481–5.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  270. Morsy SG, Tonne JM, Zhu Y, Lu B, Budzik K, Krempski JW, Ali SA, El-Feky MA, Ikeda Y. Divergent susceptibilities to AAV-SaCas9-gRNA vector-mediated genome-editing in a single-cell-derived cell population. BMC Res Notes. 2017;10(1):720.

    Article  PubMed  PubMed Central  Google Scholar 

  271. Koo T, Lu-Nguyen NB, Malerba A, Kim E, Kim D, Cappellari O, Cho HY, Dickson G, Popplewell L, Kim JS. Functional rescue of dystrophin deficiency in mice caused by frameshift mutations using Campylobacter jejuni Cas9. Mol Ther. 2018;26(6):1529–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  272. Fujii W, Ito H, Kanke T, Ikeda A, Sugiura K, Naito K. Generation of genetically modified mice using SpCas9-NG engineered nuclease. Sci Rep. 2019;9(1):12878.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  273. Casini A, Olivieri M, Petris G, Montagna C, Reginato G, Maule G, Lorenzin F, Prandi D, Romanel A, Demichelis F, et al. A highly specific SpCas9 variant is identified by in vivo screening in yeast. Nat Biotechnol. 2018;36(3):265–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  274. Hu JH, Miller SM, Geurts MH, Tang W, Chen L, Sun N, Zeina CM, Gao X, Rees HA, Lin Z, 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 

  275. Miller SM, Wang T, Randolph PB, Arbab M, Shen MW, Huang TP, Matuszek Z, Newby GA, Rees HA, Liu DR. Continuous evolution of SpCas9 variants compatible with non-G PAMs. Nat Biotechnol. 2020;38(4):471–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  276. Hu Z, Yu L, Zhu D, Ding W, Wang X, Zhang C, Wang L, Jiang X, Shen H, He D. Disruption of HPV16-E7 by CRISPR/Cas system induces apoptosis and growth inhibition in HPV16 positive human cervical cancer cells. BioMed Res Int. 2014;2014:1.

    Article  Google Scholar 

  277. Jubair L, Lam AK, Fallaha S, McMillan NA. CRISPR/Cas9-loaded stealth liposomes effectively cleared established HPV16-driven tumours in syngeneic mice. PLoS ONE. 2021;16(1):e0223288.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  278. Ebina H, Misawa N, Kanemura Y, Koyanagi Y. Harnessing the CRISPR/Cas9 system to disrupt latent HIV-1 provirus. Sci Rep. 2013;3:2510.

    Article  PubMed  PubMed Central  Google Scholar 

  279. Li X, Guo M, Hou B, Zheng B, Wang Z, Huang M, Xu Y, Chang J, Wang T. CRISPR/Cas9 nanoeditor of double knockout large fragments of E6 and E7 oncogenes for reversing drugs resistance in cervical cancer. J Nanobiotechnol. 2021;19(1):1–13.

    Google Scholar 

  280. Lin S-R, Yang H-C, Kuo Y-T, Liu C-J, Yang T-Y, Sung K-C, Lin Y-Y, Wang H-Y, Wang C-C, Shen Y-C. The CRISPR/Cas9 system facilitates clearance of the intrahepatic HBV templates in vivo. Mol Ther Nucleic Acids 2014;3:e186.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  281. Zhen S, Hua L, Takahashi Y, Narita S, Liu Y-H, Li Y. In vitro and in vivo growth suppression of human papillomavirus 16-positive cervical cancer cells by CRISPR/Cas9. Biochem Biophys Res Commun. 2014;450(4):1422–6.

    Article  CAS  PubMed  Google Scholar 

  282. Jiang C, Mei M, Li B, Zhu X, Zu W, Tian Y, Wang Q, Guo Y, Dong Y, Tan X. A non-viral CRISPR/Cas9 delivery system for therapeutically targeting HBV DNA and pcsk9 in vivo. Cell Res. 2017;27(3):440–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  283. Liao H-K, Gu Y, Diaz A, Marlett J, Takahashi Y, Li M, Suzuki K, Xu R, Hishida T, Chang C-J. Use of the CRISPR/Cas9 system as an intracellular defense against HIV-1 infection in human cells. Nat Commun. 2015;6(1):1–10.

    Article  CAS  Google Scholar 

  284. Yang Y-C, Chen Y-H, Kao J-H, Ching C, Liu I-J, Wang C-C, Tsai C-H, Wu F-Y, Liu C-J, Chen P-J. Permanent inactivation of HBV genomes by CRISPR/Cas9-mediated non-cleavage base editing. Mol Ther Nucleic Acids 2020;20:480–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  285. Gao X, Jin Z, Tan X, Zhang C, Zou C, Zhang W, Ding J, Das BC, Severinov K, Hitzeroth II. Hyperbranched poly (β-amino ester) based polyplex nanopaticles for delivery of CRISPR/Cas9 system and treatment of HPV infection associated cervical cancer. J Control Release. 2020;321:654–68.

    Article  CAS  PubMed  Google Scholar 

  286. Kaminski R, Chen Y, Fischer T, Tedaldi E, Napoli A, Zhang Y, Karn J, Hu W, Khalili K. Elimination of HIV-1 genomes from human T-lymphoid cells by CRISPR/Cas9 gene editing. Sci Rep. 2016;6(1):1–15.

    CAS  Google Scholar 

  287. Wang W, Ye C, Liu J, Zhang D, Kimata JT, Zhou P. CCR5 gene disruption via lentiviral vectors expressing Cas9 and single guided RNA renders cells resistant to HIV-1 infection. PLoS ONE. 2014;9(12):e115987.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  288. Inturi R, Jemth P. CRISPR/Cas9-based inactivation of human papillomavirus oncogenes E6 or E7 induces senescence in cervical cancer cells. Virology. 2021;562:92–102.

    Article  CAS  PubMed  Google Scholar 

  289. Hsu DS, Kornepati AV, Glover W, Kennedy EM, Cullen BR. Targeting HPV16 DNA using CRISPR/Cas inhibits anal cancer growth in vivo. Futur Virol. 2018;13(07):475–82.

    Article  CAS  Google Scholar 

  290. Yu L, Hu Z, Gao C, Feng B, Wang L, Tian X, Ding W, Jin X, Ma D, Wang H. deletion of HPV18 E6 and E7 genes using dual sgRNA-directed CRISPR/Cas9 inhibits growth of cervical cancer cells. Int J Clin Exp Med. 2017;10(6):9206–13.

    Google Scholar 

  291. Kaushik A, Yndart A, Atluri V, Tiwari S, Tomitaka A, Gupta P, Jayant RD, Alvarez-Carbonell D, Khalili K, Nair M. Magnetically guided non-invasive CRISPR-Cas9/gRNA delivery across blood-brain barrier to eradicate latent HIV-1 infection. Sci Rep. 2019;9(1):1–11.

    Article  Google Scholar 

  292. Lao YH, Li M, Gao MA, Shao D, Chi CW, Huang D, Chakraborty S, Ho TC, Jiang W, Wang HX. HPV oncogene manipulation using nonvirally delivered CRISPR/Cas9 or Natronobacterium gregoryi Argonaute. Adv Sci 2018;5(7):1700540.

    Article  CAS  Google Scholar 

  293. Ye L, Wang J, Beyer AI, Teque F, Cradick TJ, Qi Z, Chang JC, Bao G, Muench MO, Yu J. Seamless modification of wild-type induced pluripotent stem cells to the natural CCR5Δ32 mutation confers resistance to HIV infection. Proc Natl Acad Sci. 2014;111(26):9591–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  294. Nguyen H, Wilson H, Jayakumar S, Kulkarni V, Kulkarni S. Efficient inhibition of HIV using CRISPR/Cas13d nuclease system. Viruses. 2021;13(9):1850.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  295. Liu S, Wang Q, Yu X, Li Y, Guo Y, Liu Z, Sun F, Hou W, Li C, Wu L. HIV-1 inhibition in cells with CXCR4 mutant genome created by CRISPR-Cas9 and piggyBac recombinant technologies. Sci Rep. 2018;8(1):1–11.

    Google Scholar 

  296. Yoshiba T, Saga Y, Urabe M, Uchibor R, Matsubara S, Fujiwara H, Mizukami H. CRISPR/Cas9-mediated cervical cancer treatment targeting human papillomavirus E6. Oncol Lett. 2019;17(2):2197–206.

    CAS  PubMed  Google Scholar 

  297. Wei L, Chiu DK, Tsang FH, Law CT, Cheng CL, Au SL, Lee JM, Wong CC, Ng IO, Wong CM. Histone methyltransferase G9a promotes liver cancer development by epigenetic silencing of tumor suppressor gene RARRES3. J Hepatol. 2017;67(4):758–69.

    Article  CAS  PubMed  Google Scholar 

  298. Song J, Zhang X, Ge Q, Yuan C, Chu L, Liang HF, Liao Z, Liu Q, Zhang Z, Zhang B. CRISPR/Cas9-mediated knockout of HBsAg inhibits proliferation and tumorigenicity of HBV-positive hepatocellular carcinoma cells. J Cell Biochem. 2018;119(10):8419–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  299. Dai C, Zhang X, Xie D, Tang P, Li C, Zuo Y, Jiang B, Xue C. Targeting PP2A activates AMPK signaling to inhibit colorectal cancer cells. Oncotarget. 2017;8(56):95810–23.

    Article  PubMed  PubMed Central  Google Scholar 

  300. Hannafon BN, Cai A, Calloway CL, Xu Y-F, Zhang R, Fung K-M, Ding W-Q. miR-23b and miR-27b are oncogenic microRNAs in breast cancer: evidence from a CRISPR/Cas9 deletion study. BMC Cancer. 2019;19(1):642.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  301. Singhal J, Chikara S, Horne D, Awasthi S, Salgia R, Singhal SS. Targeting RLIP with CRISPR/Cas9 controls tumor growth. Carcinogenesis. 2021;42(1):48–57.

    Article  CAS  PubMed  Google Scholar 

  302. Zhang S, Fan G, Hao Y, Hammell M, Wilkinson JE, Tonks NK. Suppression of protein tyrosine phosphatase N23 predisposes to breast tumorigenesis via activation of FYN kinase. Genes Dev. 2017;31(19):1939–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  303. McDermott MS, Chumanevich AA, Lim CU, Liang J, Chen M, Altilia S, Oliver D, Rae JM, Shtutman M, Kiaris H, et al. Inhibition of CDK8 mediator kinase suppresses estrogen dependent transcription and the growth of estrogen receptor positive breast cancer. Oncotarget. 2017;8(8):12558–75.

    Article  PubMed  PubMed Central  Google Scholar 

  304. Zheng JJ, He Y, Liu Y, Li FS, Cui Z, Du XM, Wang CP, Wu YM. Novel role of PAF1 in attenuating radiosensitivity in cervical cancer by inhibiting IER5 transcription. Radiat Oncol. 2020;15(1):131.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  305. Gao G, Zhang L, Villarreal OD, He W, Su D, Bedford E, Moh P, Shen J, Shi X, Bedford MT, et al. PRMT1 loss sensitizes cells to PRMT5 inhibition. Nucleic Acids Res. 2019;47(10):5038–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  306. Jiang L, Chen T, Xiong L, Xu JH, Gong AY, Dai B, Wu G, Zhu K, Lu E, Mathy NW, et al. Knockdown of m6A methyltransferase METTL3 in gastric cancer cells results in suppression of cell proliferation. Oncol Lett. 2020;20(3):2191–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  307. Jin J, Xie S, Sun Q, Huang Z, Chen K, Guo D, Rao X, Deng Y, Liu Y, Li S, et al. Upregulation of BCAM and its sense lncRNA BAN are associated with gastric cancer metastasis and poor prognosis. Mol Oncol. 2020;14(4):829–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  308. Muniyan S, Haridas D, Chugh S, Rachagani S, Lakshmanan I, Gupta S, Seshacharyulu P, Smith LM, Ponnusamy MP, Batra SK. MUC16 contributes to the metastasis of pancreatic ductal adenocarcinoma through focal adhesion mediated signaling mechanism. Genes Cancer. 2016;7(3–4):110–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  309. Chen L, Alexe G, Dharia NV, Ross L, Iniguez AB, Conway AS, Wang EJ, Veschi V, Lam N, Qi J, et al. CRISPR-Cas9 screen reveals a MYCN-amplified neuroblastoma dependency on EZH2. J Clin Invest. 2018;128(1):446–62.

    Article  PubMed  Google Scholar 

  310. Rodriguez AC, Vahrenkamp JM, Berrett KC, Clark KA, Guillen KP, Scherer SD, Yang CH, Welm BE, Janát-Amsbury MM, Graves BJ, et al. ETV4 is necessary for estrogen signaling and growth in endometrial cancer cells. Cancer Res. 2020;80(6):1234–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  311. Bian X, Gao J, Luo F, Rui C, Zheng T, Wang D, Wang Y, Roberts TM, Liu P, Zhao JJ, et al. PTEN deficiency sensitizes endometrioid endometrial cancer to compound PARP-PI3K inhibition but not PARP inhibition as monotherapy. Oncogene. 2018;37(3):341–51.

    Article  CAS  PubMed  Google Scholar 

  312. Prattapong P, Ngernsombat C, Aimjongjun S, Janvilisri T. CRISPR/Cas9-mediated double knockout of SRPK1 and SRPK2 in a nasopharyngeal carcinoma cell line. Cancer Rep (Hoboken). 2020;3(2):e1224.

    CAS  PubMed  Google Scholar 

  313. Liu H, Li Z, Huo S, Wei Q, Ge L. Induction of G0/G1 phase arrest and apoptosis by CRISPR/Cas9-mediated knockout of CDK2 in A375 melanocytes. Mol Clin Oncol. 2020;12(1):9–14.

    CAS  PubMed  Google Scholar 

  314. Zhao Q, Qian Q, Cao D, Yang J, Gui T, Shen K. Role of BMI1 in epithelial ovarian cancer: investigated via the CRISPR/Cas9 system and RNA sequencing. J Ovarian Res. 2018;11(1):31.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  315. Zhen S, Hua L, Liu YH, Sun XM, Jiang MM, Chen W, Zhao L, Li X. Inhibition of long non-coding RNA UCA1 by CRISPR/Cas9 attenuated malignant phenotypes of bladder cancer. Oncotarget. 2017;8(6):9634–46.

    Article  PubMed  Google Scholar 

  316. Huang LC, Tam KW, Liu WN, Lin CY, Hsu KW, Hsieh WS, Chi WM, Lee AW, Yang JM, Lin CL, et al. CRISPR/Cas9 genome editing of epidermal growth factor receptor sufficiently abolished oncogenicity in anaplastic thyroid cancer. Dis Markers. 2018;2018:3835783.

    Article  PubMed  PubMed Central  Google Scholar 

  317. Kailayangiri S, Altvater B, Lesch S, Balbach S, Göttlich C, Kühnemundt J, Mikesch JH, Schelhaas S, Jamitzky S, Meltzer J, et al. EZH2 inhibition in Ewing sarcoma upregulates G(D2) expression for targeting with gene-modified T cells. Mol Ther. 2019;27(5):933–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  318. Cooper ML, Choi J, Staser K, Ritchey JK, Devenport JM, Eckardt K, Rettig MP, Wang B, Eissenberg LG, Ghobadi A, et al. An “off-the-shelf” fratricide-resistant CAR-T for the treatment of T cell hematologic malignancies. Leukemia. 2018;32(9):1970–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  319. Jung IY, Kim YY, Yu HS, Lee M, Kim S, Lee J. CRISPR/Cas9-mediated knockout of DGK improves antitumor activities of human T cells. Cancer Res. 2018;78(16):4692–703.

    Article  CAS  PubMed  Google Scholar 

  320. Dai X, Park JJ, Du Y, Kim HR, Wang G, Errami Y, Chen S. One-step generation of modular CAR-T cells with AAV-Cpf1. Nat Methods. 2019;16(3):247–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  321. Ren J, Zhang X, Liu X, Fang C, Jiang S, June CH, Zhao Y. A versatile system for rapid multiplex genome-edited CAR T cell generation. Oncotarget. 2017;8(10):17002–11.

    Article  PubMed  PubMed Central  Google Scholar 

  322. Choi BD, Yu X, Castano AP, Darr H, Henderson DB, Bouffard AA, Larson RC, Scarfò I, Bailey SR, Gerhard GM, et al. CRISPR-Cas9 disruption of PD-1 enhances activity of universal EGFRvIII CAR T cells in a preclinical model of human glioblastoma. J Immunother Cancer. 2019;7(1):304.

    Article  PubMed  PubMed Central  Google Scholar 

  323. Stenger D, Stief TA, Kaeuferle T, Willier S, Rataj F, Schober K, Vick B, Lotfi R, Wagner B, Grünewald TGP, et al. Endogenous TCR promotes in vivo persistence of CD19-CAR-T cells compared to a CRISPR/Cas9-mediated TCR knockout CAR. Blood. 2020;136(12):1407–18.

    Article  PubMed  Google Scholar 

  324. Giuffrida L, Sek K, Henderson MA, Lai J, Chen AXY, Meyran D, Todd KL, Petley EV, Mardiana S, Mølck C, et al. CRISPR/Cas9 mediated deletion of the adenosine A2A receptor enhances CAR T cell efficacy. Nat Commun. 2021;12(1):3236.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

No funders.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the conception and the main idea of the work. S.S.B., Z.R.G., B.A.Z., S.A.H.A., A.İ., L.T., N.S., and A.A. drafted the main text, figures, and tables. R.M. and A.F.Y. equally supervised the work and provided comments and additional scientific information. S.S.B. and Z.R.G. also reviewed and revised the text. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Roozbeh Moghaddar or Amirhossein Fakhre Yaseri.

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.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

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/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shojaei Baghini, S., Gardanova, Z.R., Abadi, S.A.H. et al. CRISPR/Cas9 application in cancer therapy: a pioneering genome editing tool. Cell Mol Biol Lett 27, 35 (2022). https://doi.org/10.1186/s11658-022-00336-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s11658-022-00336-6

Keywords