Abstract
Cancer is a genetic disease caused by alterations in genome and epigenome and is one of the leading causes for death worldwide. The exploration of disease development and therapeutic strategies at the genetic level have become the key to the treatment of cancer and other genetic diseases. The functional analysis of genes and mutations has been slow and laborious. Therefore, there is an urgent need for alternative approaches to improve the current status of cancer research. Gene editing technologies provide technical support for efficient gene disruption and modification in vivo and in vitro, in particular the use of clustered regularly interspaced short palindromic repeats (CRISPR)–Cas systems. Currently, the applications of CRISPR–Cas systems in cancer rely on different Cas effector proteins and the design of guide RNAs. Furthermore, effective vector delivery must be met for the CRISPR–Cas systems to enter human clinical trials. In this review article, we describe the mechanism of the CRISPR–Cas systems and highlight the applications of class II Cas effector proteins. We also propose a synthetic biology approach to modify the CRISPR–Cas systems, and summarize various delivery approaches facilitating the clinical application of the CRISPR–Cas systems. By modifying the CRISPR–Cas system and optimizing its in vivo delivery, promising and effective treatments for cancers using the CRISPR–Cas system are emerging.
Keywords: cancer, CRISPR–Cas, gene editing, synthetic biology, vector delivery
Abbreviations
- AML
acute myelogenous leukemia
- AAVs
adeno‐associated viruses
- AVs
adenoviruses
- BP
black phosphorus
- CPPs
cell‐penetrating peptides
- CRISPR
clustered regularly interspaced short palindromic repeats
- crRNAs
CRISPR RNAs
- CTLs
cytotoxic T lymphocytes
- DSB
double‐strand breaks
- EVs
extracellular vesicles
- FRL
far‐infrared light
- AuNPs
gold nanoparticles
- GO
graphene oxide
- HSC
hematopoietic stem cell
- hgRNA
homing guide RNA
- HDR
homology‐directed repair
- hTERT
human telomerase reverse transcriptase
- hUP II
human uroplakin II gene
- LVs
lentiviruses
- LAN
lance array nanoinjection
- NIR
near‐infrared light
- NHEJ
nonhomologous end joining
- PC
pancreatic cancer
- PDAC
pancreatic ductal adenocarcinoma
- PEI
polyethyleneimine
- PAM
protospacer adjacent motif
- PFS
protospacer flanking site
- RNP
ribonucleoprotein
- scRNA
scaffold RNA
- shRNA
short hairpin RNA
- SNMs
single nucleotide mutations
- sgRNA
single‐stranded guide RNA
- SMA
spinal muscular atrophy
- ThT
thiaxanthin T
- TALENs
transcription activator‐like effector nucleases
- UCNPs
upconversion nanoparticles
- ZFNs
zinc‐finger nucleases
1. INTRODUCTION
Cancer is one of the most important health problems worldwide. It was estimated that there were 19.3 million new cases and 10 million cancer deaths worldwide in 2020. 1 Cancer remains the leading disease‐associated cause of death, despite rapid advances in treatments. 2 Traditional therapies suffer from disadvantages such as poor specificity and resistance to chemotherapy drugs. 3 The exact mechanisms underlying the poor effectiveness of traditional therapies are not yet clear. Thus, uncovering the genetic pattern of cancer and subsequently deepening the understanding of its role in cancer development and response to treatment has become a major focus of attention. 4
The availability of zinc‐finger nucleases (ZFNs) and transcription activator‐like effector nucleases (TALENs) offers the possibility of directly targeting and modifying genomic sequences. 5 , 6 However, ZFNs and TALENs are expensive and inefficient, limiting their clinical application. 7 In this aspect, clustered regularly interspaced short palindromic repeats and CRISPR‐associated genes (CRISPR–Cas) system offers promising solutions to such limitations, with multiple advantages. 8 Yan et al. 9 used CRISPR/Cas9 to knock down miR‐3064 and showed that the proliferation, invasion, and tumorigenic ability of pancreatic cancer (PC) cells were significantly inhibited. Meanwhile, CRISPR–Cas system could define vulnerabilities in cancer by identifying essential genes, gene interactions, and anticancer immune targets. 10 Importantly, the advent of multiple toolkits further extended the application of the CRISPR–Cas systems in cancer.
In this review, we conducted an in‐depth analysis of the mechanism of action of the CRISPR–Cas systems, and the research progress of newly emerging toolboxes to modify and deliver CRISPR systems. We also highlighted the potential added effects of incorporating synthetic biology into the CRISPR–Cas systems for cancer treatment. We further explored the delivery issues currently faced by the CRISPR–Cas system for clinical studies, with the hope of further refining the CRISPR–Cas systems to facilitate clinical application.
2. THE COMPONENTS AND MECHANISMS OF CRISPR–Cas SYSTEMS
The CRISPR–Cas systems, consisting of CRISPR arrays and highly diverse Cas genes, are adaptive immune systems evolved by bacteria and archaea in their immune system against invading phages and foreign plasmid DNA 11 , 12 (Figure 1). Structurally, CRISPR arrays contain a leader (adjoining the first repeat of CRISPR loci and considered as the promoter of CRISPR arrays), short direct repeats (forming hairpin structures to stabilize the secondary structure of RNA), and nonrepetitive spacers (captured exogenous DNA sequences) 13 , 14 (Figure 2a). These arrays can be transcribed and processed into CRISPR RNAs (crRNAs), which are used to direct Cas nucleases to cleave complementary exogenous DNA sequences. 15 At least 45 natural Cas proteins have been identified in different bacteria, as exemplified by the well‐known Streptococcus thermophilus (St1) has three Cas genes: Cas9, Cas1, and Cas2. 16 The domain organization of SpCas9 consists of NUC lobe and REC lobe (Figure 2b). Meanwhile, depending on the architecture of the CRISPR array and the signature interference effector, CRISPR–Cas systems can be classified into two classes [containing six types (I–VI) and 33 sub‐types]. 17 , 18 Class I systems (including type I, III, and IV) encompass multisubunit Cas effector proteins, which bind to crRNA and generate target interference. Additionally, class II systems (including type II, V, and VI) require only a single, multidomain large Cas effector protein to form a complex with crRNA in the interference process. 19 Accordingly, the class II systems represented by CRISPR–Cas9 require only one Cas effector protein to function as a cleavage, 20 while the class I systems demand multiple Cas effector proteins, limiting their applications. 21 Hence, class II systems exhibit tremendous promises for genome engineering in cleavage of target DNA and RNA.
Simply put, bacteria and archaea are able to store a small segment of viral gene (named spacer) into the CRISPR array when they are first invaded by a virus. When the same virus invades again, the bacteria are able to recognize the virus based on the spacer and disable it by cutting off the DNA of the virus. The specific process involves three major steps (using CRISPR–Cas9 as an example): adaptation, expression, and interference. The adaptation stage is the spacer acquisition to form memory of previous infections and is what makes CRISPR–Cas immunization adaptive and heritable. Spacer acquisition relies on Cas1 and Cas2 which are present in almost all CRISPR–Cas systems. 18 Cas1 is catalytic and Cas2 has a structural function. 22 Adaptation mechanisms show a preference for foreign DNA over self‐DNA as the key to avoiding autoimmunity. For example, the RecBCD repair complex present in Escherichia coli is able to degrade larger portions of the foreign genome and serves as the basis for preferential access to nonself DNA. 23 Once the exogenous DNA is injected into the host, the proteins encoded by Cas1 and Cas2 would recognize protospacer adjacent motif (PAM) in the exogenous DNA sequence, and then take DNA sequence adjacent to the PAM as a protospacer. 24 Next, the Cas1/2 protein complex snips the protospacer from exogenous DNA to form a spacer which is inserted between two repeats at the 5′‐end of CRRSPR array with the assistance of enzymes. 25 , 26
At the expression stage, CRISPR array (repeats and spacers) is transcribed to generate pre‐crRNA (crRNA precursor) and tracrRNA (trans‐activating CRISPR RNA) under the initiation of leader. Pre‐crRNA is sheared into mature crRNA (containing 1 spacer and several repeats) by RNase III nuclease and Cas9 protein. Repeats and tracrRNA form a double‐stranded RNA (tracrRNA‐crRNA) by base complementary pairing which assembles into a complex with Cas9 protein. 27 The effector complex composed of Cas9 and the tracrRNA:crRNA duplex exerts interference after a second cleavage by an unknown RNase—which removes the 5′ repeat‐derived tag. 28 The spacer is in free state. This complex would monitor exogenous DNA sequence at all times. 29
During the interference stage, the Cas9 effector protein is already bound to the guide RNA prior to target selection and cleavage, thus participating in crRNA maturation. The spacer recognizes the complementary sequence in exogenous DNA. The entire complex is also localized to specific PAM, and the DNA double‐strand is then unraveled. The spacer sequence hybridizes to the complementary strand, while the other strand remains free. Subsequently, Cas9 protein is localized to the correct PAM sequence. 30 Base pairing of crRNA with the target strand induces an R‐loop structure that eventually triggers cleavage of the target and nontarget strands by the domains of Cas9 protein, respectively, resulting in flat‐end cleavage at three nucleotides upstream of the PAM. 28 This eventually results in DNA double‐strand breaks (DSB), silencing of exogenous DNA expression, and successful immunization. 31 Thus, the interference protects hosts from invasion of exogenous genome and also gives a chance for gene editing since typical CRISPR–Cas9 system brings a break of double‐stranded DNA.
It can be seen that CRISPR array is used to identify exogenous DNA sequences, while Cas9 protein acts as scissors for cleavage. Among them, tracrRNA–crRNA serves as the navigator of the system and guides Cas9 for precise targeting. Importantly, in 2012, Jinek et al. 32 designed a single‐stranded guide RNA (sgRNA) that replaced the crRNA–tracrRNA complex, which can direct the Cas9 protein to specifically cleave the target gene (Figure 2c). This success confirms the feasibility of artificially designed sgRNAs for target sequence by means of synthetic biology, thus enabling gene repair and modeling of mutations, knock‐out, knock‐in, fusion. 33 Especially in recent years, the rise of multiple Cas proteins has once again brought the advantages of CRISPR–Cas systems to the forefront with remarkable achievements.
3. COMMONLY USED CAS PROTEINS AND THEIR NOVEL DERIVATIVES
Although many CRISPR–Cas systems have been identified, only a few of them have been used as research tools. Among them, class II systems relying on a single‐effector Cas protein are widely used for gene editing in mammals. Specifically, the representative effector proteins in the class II are Cas9 (type II), Cas12 (type V), and Cas13 (type VI) 34 , 35 , 36 (Figure 2d). The characteristics of the different Cas proteins can be seen in Table 1.
TABLE 1.
Type/class | Effector | Nuclease domains | Function | Size (amine acids) | Molecular weight (kDa) | PAM sequence | Length of guiding sequence (bp) | Cutting site |
---|---|---|---|---|---|---|---|---|
II/2 | spCas9 | RuvC, HNH | DNA nuclease | 1368 | 158.3 | NGG | 20 | ~3 bp 5′ of PAM |
FnCas9 | RuvC, HNH | DNA nuclease | 1628 | 190.3 | NGG | 20 | ~3 bp 5′ of PAM | |
SaCas9 | RuvC, HNH | DNA nuclease | 1052 | 123.8 | NNGR R | 21 | ~3 bp 5′ of PAM | |
NmCas9 | RuvC, HNH | DNA nuclease | 1081 | 124.2 | NNNNG ATT | 24 | ~3 bp 5′ of PAM | |
St1Cas9 | RuvC, HNH | DNA nuclease | 1122 | 129.5 | NNAGA AW | 20 | ~3 bp 5′ of PAM | |
St3Cas9 | RuvC, HNH | DNA nuclease | 1409 | 164.9 | NGGNG | 20 | ~3 bp 5′ of PAM | |
CjCas9 | RuvC, HNH | DNA nuclease | 983 | 114.8 | NNNNACAC | 22 | ~3 bp 5′ of PAM | |
V/2 | AsCpf1 | RuvC | crRNA processing, DNA nuclease | 1307 | 151.2 | TTTV | 24 | 19/24 bp 3′ of PAM |
LbCpf1 | RuvC | crRNA processing, DNA nuclease | 1228 | 143.7 | TTTV | 24 | 19/24 bp 3′ of PAM | |
VI/2 | Cas13 | HEPN × 2 | crRNA processing, RNA nuclease | Multiple orthologs | Variable | RNA targeting | 28 | / |
3.1. Cas9
Cas9, the most widely used effector protein, has two structural domains with cleavage activity: the HNH domain (responsible for cleaving the complementary DNA strand with crRNA) and the Ruvc domain (responsible for cleaving the noncomplementary DNA strand). 37 The commonly used Cas9 protein is derived from Streptococcus pyogenes (SpCas9), whose PAM sequence is NGG (N is any nucleotide). To further extend the diversity of PAM sequences, over 10 different Cas proteins have been identified in the last few years. For example, the smallest Cas9 nuclease is from Campylobacter jejuni (CjCas9), which has only 984 amino acids and its PAM sequence is NNNNNACAC. The small size of CjCas9 allows for good intracellular delivery, but its targeting range and flexibility are relatively limited. 38
Cas9 nuclease cleaves DNA and produces genome editing effects via nonhomologous end joining (NHEJ) or homology‐directed repair (HDR) pathways (Figure 2e). A variety of Cas9 mutation systems have been developed to address the off‐target effects of CRISPR–Cas9. In 2013, Ran et al. mutated one of the two catalytically active domains of Cas9 (HNH or RuvC) to obtain Cas9 nickase (Cas9n). 39 This form of mutation produces single‐stranded gaps rather than double‐stranded breaks, and allows gene editing via the HDR pathway. If double‐stranded DNA needs to be cleaved, the two gRNAs will be designed to be on opposite DNA strands and in close proximity (sequences no more than 20 bp apart), thus effectively introducing a DSB. Ultimately, on‐target stringency can be increased while off‐target mutations are minimized with Cas9n. 40 Further, by mutating both nuclease active regions of Cas9, a dead Cas9 (dCas9) that specifically recognizes only sgRNA and has no shearase activity was generated. 41 dCas9 is mainly fused with transcriptional regulatory elements or chromosomal modification elements to build new tools for the regulation of transcriptional and epigenetic modifications such as CRISPRa (transcriptional activation), CRISPRi (transcriptional interference), and CRISPRoff (controlling gene expression with high specificity while leaving the DNA unchanged). 42 , 43 , 44 For example, dCas9 promotes or represses the transcription of target genes by binding to activation domains (VP16, VP64, NF‐κB) or repression domains (KRAB, MIX1).
To further overcome the targeting limitations of PAM, Walton et al. 45 designed new Cas9 variants, named as SpG and SpRY that bind and cleave DNA without specific PAM and are capable of unrestricted targeting the majority of the human genome with single base‐pair precision. It is thus clear that the off‐target and PAM sequence defects of the CRISPR–Cas9 system can be refined, leaving the efficient delivery of CRISPR–Cas9 as the remaining obstacle for in vivo application of the CRISPR–Cas9 systems, which will be addressed in a subsequent section. Overall, optimized CRISPR–Cas9 system can be applied to a wider range of fields, including gene therapy for cancer.
3.2. Cas12
Unlike Cas9, Cas12 nuclease contains only a RuvC‐like domain that cleaves two strands to induce DSB. 46 Since possessing RNAase and DNAase activity, the Cas12 nuclease relies on a single crRNA guide for DNA localization and cleaves at the distal PAM end to produce 5‐nt sticky ends, in contrast to Cas9 which normally cleaves near the PAM end to produce blunt ends. 47 Widely used Cas12a (known as Cpf1) is from Acidaminococcus spp. (AsCas12a) and Lachnospiraceae spp. (LbCas12a), with a small molecular mass of 1200 to 1300 amino acids. 48 On the one hand, unlike the G‐rich PAMs required for Cas9, Cas12a can recognize T‐rich PAMs, thus further increasing the number of potential target sites. On the other hand, Cas12a can follow its own cleavage pattern and PAM sequences to generate staggered ends, facilitating precisely targeted integration of DNA. The restricted recognition of PAM (5′‐TTTN‐3′) by AsCas12a and LbCas12a limits their application in the field of gene editing. 49 An enhanced AsCas12a variant (enAsCas12a) has been designed to improve genome editing activity. Meanwhile, the targeting range for Cas12a has been expanded greatly by newly engineered AsCas12a variants that recognize PAMs 5′‐TYCV and 5′‐TATV, or PAMs 5′‐VTTV, 5′‐TTTT, 5′‐TTCN, and 5′‐TATV. 50 , 51 Another reported Cas12a with a Francisella novicida origin (FnCas12a) has a PAM sequence of 5′‐KYTV‐3′ (K is T and G; Y is C and T; V is A, C, and G) and possesses DNA cleavage activity in human cells at multiple loci. 52 The extended PAM sequences enhance selectable regions of target sites and enrich applications of Cas12a in gene editing.
Cas12a is not only flexible but also shows a high degree of specificity. Kim et al. 53 used Digenome‐seq to analyze the whole genome after the action of different gene‐editing enzymes. It was found that for the same crRNA, LbCas12a and AsCas12a had 6 and 12 off‐target sites, respectively, which were far fewer than those caused by Cas9 (>90 sites). Moreover, Kleinstiver et al. 54 compared the off‐targets of AsCas12a and LbCas12a with those of SpCas9 in vivo using GUIDE‐seq analysis. The results showed that Cas12a had a mutation frequency as low as ~0.1%–0.2% in the majority of off‐target sites, indicating that Cas12a has very limited off‐targets compared to SpCas9. Similar to Cas9, by mutating the RuvC domain, a catalytically inactive version of Cas12a (dCas12a) was generated. 55 dCas12a is able to bind different enzymes, thereby mediating transcriptional activation. For example, potent transcriptional activation of dLbCas12a can be achieved by fusing synthetic activation complexes consisting of VP16, p65, and the Rta activator domains. 56 The SunTag system is known to contain multiple scFv fusions to VP64, where scFv is a single‐chain variable fragment antibody against GCN4. 57 The dLbCas12a gene can be efficiently activated by binding SunTag to the C‐terminus of dLbCas12a. 58 However, transcriptional activation was only observed with dLbCas12a, whereas dAsCas12a induced merely marginal activation. 58
The smaller size of Cas12a facilitates intracellular delivery. Also, PAM distal cleavage ensures the safety of target recognition of sgRNAs by the small indels generated through the NHEJ repair pathway. Importantly, Cas12a has been demonstrated to have high specificity and low off‐target effects. Consequently, emerging new Cas derivatives including Cas12b, Cas12c, Cas12d (CasY), Cas12e (CasX), Cas12g, Cas12h, Cas12i, and Cas12j (CasΦ), 59 , 60 , 61 , 62 will greatly expand gene‐editing adaptations in tumor therapy.
3.3. Cas13
Both Cas9 and Cas12a have been shown to target DNA, but no single Cas effector protein can target RNA. In 2016, Shmakov et al. 20 identified a new Cas effector protein named as C2c2 (Cas13a) that can target RNA. Subsequently, Liu et al. 63 , 64 and Knott et al. 65 reported that Cas13a contains two HEPN domains with RNase activity. Since then, another three Cas13 family proteins were identified, namely Cas13b, Cas13c, and Cas13d. 66 , 67 Cas13a is a two‐component system with the binding domain of the catalytic site on the outer surface of the protein, which performs cis or trans cleavage of RNA. 64 The Cas13b system lacks Cas1 and Cas2 and is capable of targeting the secondary structure of RNA. 68 Cas13d is 20%–30% smaller than other Cas13 subtypes, facilitating its flexible packaging and intracellular delivery. 69 There is very limited information for Cas13c system.
Unlike Cas9, Cas13 proteins do not require PAM sequences to identify their targets, but they do have a protospacer flanking site (PFS) structure dependency, that is, that the base before the original spacer sequence should be A, C, or U. 70 , 71 , 72 Notably, in eukaryotic and prokaryotic cells, Cas13 is activated after recognition and cleavage of RNA, but also has “collateral shearing” RNase activity, which can shear adjacent single‐stranded RNA and cause cell dormancy or programmed cell death. 70 , 73 Simultaneously, dCas13 was generated by inactivating the catalytically active domain of Cas13. A stronger inhibitory effect was observed with CasRx, an engineered variant of Cas13d, exerting a lower and virtually absent off‐target effect compared to CRISPRi (i.e., dCas9) as well as short hairpin RNA (shRNA) knockdown. 74 In addition, inducible CRISPR artificial shearing factors (iCASFx) and multiplexing with orthogonal dCas13s were engineered, and their therapeutic potential was demonstrated in spinal muscular atrophy (SMA) fibroblasts. 75 Altogether, the nonspecific RNase activity of Cas13 provides powerful tools for applications in nucleic acid detection, gene regulation, RNA imaging, and cancer therapy. 76 Particularly, Cas13 has great potential for cancer therapy by targeted manipulation of key cancer‐associated RNA molecules (including mRNA and ncRNAs). 77 , 78
4. COMBINED APPLICATIONS OF CRISPR WITH SYNTHETIC BIOLOGY IN CANCER
Besides its naturally occurring form, the CRISPR–Cas system can also be modified by other means. Synthetic biology is a combination of engineering and biology disciplines that uses engineering design concepts to modify or create artificial living systems, often in a way that rewires naturally occurring biological circuits (either genes or proteins) to achieve the desired logical forms of cellular control. 79 In the last decade, synthetic biology has begun to develop rapidly, with a series of pioneering milestones such as the smallest artificial synthetic cell (named JCVI‐Syn3.0), 80 and artificial cells that can grow and divide normally. 81 Gradually, we could “read,” “write,” and “compile” the genome, and have the ability to design and synthesize life. Today, biological design by applying engineering principles such as standardization, 82 modularity, 83 digital logic, 84 and mathematically predictable behavior 85 has become central to synthetic biology.
Notably, synthetic biology is developing rapidly in the field of medicine and is bound to have a dramatic impact on the medical field. 86 For example, Williams et al. 87 constructed a diverse library of multi‐receptor cell–cell recognition circuits by using synthetic Notch receptors to allow engineered T cells to achieve precise recognition of extra‐ and intracellular antigens through AND/OR/NOT gates. This approach opened up new avenues for precise identification of cells and was expected to be used to regulate the expression of target genes and kill target cells. Importantly, the application of combinatorial optimization strategies in synthetic biology, such as DNA barcoding tools and high‐throughput screening biosensors, allows us to achieve a best outcome. 88 In parallel, synthetic biology is showing great fascination in the medical field as new developments in biotechnology are expanded into its toolkit.
The construction of genetic engineering at the cell level by CRISPR‐based technologies has greatly contributed to our understanding and controlling mammalian biological systems. The lack of effective gene activators makes it difficult to regulate gene expression efficiently. Dong et al. 89 developed a synthetic bacterial transcriptional activator in E. coli by linking the activation domain to a programmable CRISPR–Cas DNA binding domain. The entire gene expression program can be turned on by inducing expression of the CRISPR–Cas systems. Thus, manipulating CRISPR technologies with engineering principles enables an efficient regulation of gene expression and provides the basis for engineering synthetic bacterial cellular devices. Meanwhile, the successful application of CRISPR systems armed with synthetic biology underscored their robust potentials in medicine.
4.1. Genetic circuits in cancer
The robust and precise switching on and off one or more genes of interest is essential for many biological circuits as well as for industrial applications. Synthetic biology aims at designing modular genetic circuits. At present, CRISPR‐based genetic circuits include logic gates, cascades, bistable switches, and temporal and spatial pattern generators, among which logic gates are the most widely explored 90 (Figure 3a).
Initially, CRISPR systems were used for the design and application of individual logic gates. The dCas9‐Mxi1‐based NOR gate designed in Saccharomyces cerevisiae enables direct conversion of gRNA inputs into gRNA outputs, allowing the gates to be “wired” together. It implements arbitrary internal logic for a variety of synthetic cellular decision‐making systems and shows minimal leak transcriptionally and digital responses, forming the basis for large, synthetic, cellular decision‐making systems. 91 By using scaffold RNA (scRNA), Hofmann et al. 92 designed a logic AND gate dependent on two independent switchable parts, namely dCas9 and MCP‐VP64 expression, controlled by galactose or β‐estradiol addition, respectively. A mathematical model and single‐cell analysis showed that the AND gate could only be activated in presence of both galactose and β‐estradiol, which in turn induced transcriptional activation and sensitive tunability. The scRNA CRISPR‐dCas9 platform provides an expression control system that can bind different components to an AND gate to achieve expression of multiple target genes with a tunable fold activation.
The complex intracellular signaling limits the application of individual logic gates, and the combined design of multiple logic gates demonstrates a more refined intracellular response. Peng et al. 93 first used CRISPR–Cas12a system to successfully constructed three 2‐input elementary AND, OR, and INHIBIT logic gates. These logic gates are capable of normal operation and can be used for rapid and sensitive detection of external pathogenic bacteria such as Staphylococcus aureus. On the other hand, to detect diverse types of disease‐relevant cues, Kempton et al. 94 developed a split Cas12a platform capable of spontaneously reassembling and demonstrated that it achieved construction of multi‐input, multi‐output logic circuits in mammalian cells. The platform was highly programmable and allowed expandable AND gates with two, three, and four inputs. The availability of anti‐CRISPR systems served as an OFF switch via incorporating NOT logic. In terms of output, the platform can generate activation and repression of target genes or permanent modification of genomic DNA. Thus, split dCas12a can achieve specific activation of the anti‐tumoral programs in cancer cells by coupling to multiple cancer‐relevant inputs.
Single nucleotide mutations (SNMs) are extremely important in cancer, and multiple SNMs further complicate the pathogenesis of cancer. 95 Therefore, performing multi‐SNMs detection and analysis facilitates cancer screening. The logic gate circuits based on CRISPR systems enable the detection of multiple SNMs. Long ssDNA fragments with guanine‐rich sequences were known to be generated after CRISPR–Cas9 system shear. The G‐quadruplex/ThT (G4/ThT) complex, as fluorescent probe, can be formed by introducing thiaxanthin T (ThT) into the G‐quadruplex. Complicate genetic locus could be distinguished by designing an AND logic gate system. The operation was set as the recognition of two mutation sites by sgRNAs as input signals and the fluorescence signals as output signals. Specifically, the input case was defined as (1,1) when both sgRNA1 and sgRNA2 existed. Subsequently, the G4/ThT complex was formed after Cas9n/sgRNA complex recognized the two cleavage sites. The fluorescent signal was activated and the process was treated as “ON”. The input cases (1, 0 or 0, 1 or 0,0) generated insufficient fluorescent signal, and the output was considered as “OFF.” 96 Furthermore, the human telomerase reverse transcriptase (hTERT) promoter is considered to be a cancer‐specific promoter, 97 while the human uroplakin II gene (hUP II) promoter is a bladder‐specific promoter. 98 Liu et al. 99 designed an AND gate based on CRISPR–Cas9 system by combining hTERT and hUP II. The hTERT and hUP II promoters served as input signals to drive transcription of Cas9 mRNA and sgRNA targeting LacI, respectively. The luciferase reporter was used as the output gene. The results showed that the circuit could detect bladder cancer cells specifically and effectively inhibit the growth of cancer cells by regulating related genes. Aiming to further improve the efficiency of cancer diagnosis so as to facilitate clinical translation, Liu et al. 100 improved and optimized the previous logic circuit to develop AND gate minigene circuits based on CRISPReader. The results showed that the minigene circuit had higher cancer recognition and intervention ability in vitro and stronger anti‐cancer effects in vivo than the conventional gene circuit. It can be anticipated that the minigene circuit will take the advantages of gene circuits in synthetic biology to further optimize precision medicine research in malignant tumors and other diseases. With the advantages of high programmability, modularity, and orthogonality, CRISPR‐based genetic circuits will enable us to construct more complicated and sophisticated synthetic circuits, making them the more powerful therapeutics for cancer therapy and other medical treatments.
4.2. Optogenetic devices in cancer
CRISPR systems‐mediated gene editing has been well‐established as a powerful tool for in vitro and in vivo gene regulation, but has not yet been able to achieve such regulation in spatial and temporal manners. Optogenetics offers an unprecedented ability to achieve precise spatial and temporal control of cellular activity using light of appropriate intensity and wavelength as a trigger signal. 101 , 102 Hence, incorporating optogenetics into CRISPR‐based gene regulation system could circumvent such a limitation.
Polstein et al. 103 engineered a CRISPR–Cas9 effector (LACE) system that induced endogenous gene transcription under blue light by fusing light‐inducible heterodimerizing proteins CRY2 and CIB1 to a transactivation domain and the catalytically inactive dCas9, respectively. The LACE system can be readily directed to DNA sequences and reversibly regulate endogenous genes by light, suggesting that optogenetic systems offer a greater flexibility and simplicity in targeting endogenous genes. Meanwhile, Niopek et al. 35 described optogenetic anti‐CRISPR variants named CASANOVA, which include AcrIIA4 (a potent Streptococcus pyogenes Cas9 inhibitor) and the LOV2 photosensor from Avena sativa. These artificial Acr proteins enable light‐mediated genome and epigenome editing via a precise external blue light stimulus. By co‐expressing CASANOVA and sgRNAs targeting different genomic loci in HEK293T cells, insertion/deletion (indel) mutations at all target loci were strongly light‐dependent. The potential of CASANOVA for Cas9 DNA targeting kinetics in living cells was further confirmed by irradiating cells with a 488 nm laser beam. Again, Niopek et al. turned their attention to the type II‐C Cas9 from Neisseria meningitidis (NmeCas9), which is much smaller than SpyCas9 with only 1081 amino acids. It exhibits high target specificity, probably related to its longer target recognition sequence (~24 nucleotides) and a longer PAM sequence (N4GATT for NmeCas9 versus NGG for SpyCas9). 104 A light‐dependent anti‐CRISPR protein named CASANOVA‐C3 was successfully developed for conditional activation of NmeCas9. Namely, NmeCas9 activity could be potently blocked in the dark, but activated under blue light, allowing reversible genome editing at various endogenous loci which could be precisely monitored and controlled using light signals. 105
Accordingly, Qi et al. 106 designed a blue light photosensor that can effectively induce the expression of Cas13a protein and precisely control the expression of MALAT1 which was confirmed to be an oncogene in human bladder cancer. The results showed that the MALAT1 expression level in bladder carcinoma cell lines (5637 and T24) was significantly suppressed and the malignant phenotype of bladder cancer cells was alleviated under blue light irradiation. This study provides a proof of concept for novel cancer treatment using the newly created CRISPR‐based gene editing system.
Due to their poor penetration and possible cytotoxicity, neither UV light nor blue light can be widely used for in vivo research applications and clinical translation. 107 , 108 Considering that red light could penetrate deep tissue more than 5 mm beneath skin surface, 107 the development of red light‐activating CRISPR systems using nanotechnology can further expand the application of optogenetic devices in cancer. A CRISPR–Cas9 nanosystem was generated using the cationic polymer‐coated Au nanorods (APC) and Cas9 plasmids driven by the heat‐inducible HSP70 promoter. APC not only serves as a vector for plasmid delivery, but also functions as an intracellular NIR light photothermal converter to induce Cas9 expression. Data from the in vitro and in vivo studies confirmed good sensitivity and inducibility of this new system, with minimal off‐target effects. 109 Furthermore, Pan et al. 110 reported an upconversion nanoparticles (UCNPs)‐based near‐infrared light (NIR)‐responsive CRISPR–Cas9 vector named UCNPs‐Cas9@PEI. UCNPs could convert NIR to localized UV light for cleavage of photosensitive molecules. NIR light stimulation induced the on‐demand release of CRISPR–Cas9, which enabled controlled gene editing. Serine/threonine‐protein kinase PLK‐1, also known as polo‐like kinase 1 (PLK‐1), has been implicated in many different cancers, therefore targeting PLK‐1 represents one of the novel therapeutics for cancer treatment. 111 , 112 Thus, by designing sgRNAs targeting PLK‐1, the anti‐tumor activity of this system can be explored. Indeed, the expression level of PLK‐1 in A549 cells was significantly inhibited with sgRNAs targeting PLK‐1. Meanwhile, tumor growth was also significantly arrested in xenograft nude mice model bearing A549 cells. This study provides clear evidence to confirm that the release of CRISPR–Cas9 with anti‐tumor potential from UCNPs‐Cas9@PEI can be precisely controlled by NIR for cancer therapy (Figure 3b). These nanoCRISPR systems with optical properties provide a programmable genome editing strategy capable of treating deep tissue tumors. Apart from NIR, far‐infrared light (FRL) has also been explored for cancer treatment. Yu et al. 113 designed an FRL‐activated split‐Cas9 (FAST) system, with two components, namely N‐terminal Cas9 fragment [Cas9(N)] linking with the Coh2 domain and C‐terminal Cas9 fragment [Cas9(C)] linking with the DocS domain. Coh2 and DocS are two C. thermocellum proteins that interact with high affinity. Under FRL illumination (730 nm), the FAST system was successfully assembled in HEK‐293 cells and activated for target genome editing. In a mouse xenograft tumor model, FRL‐triggered PLK‐1 oncogene editing was successfully achieved using FAST system. The system thus extended the spectrum of light energies in optogenetic toolbox, allowing for the noninvasive induction of gene editing activity in cells located in deep tissues. Taken together, the optogenetic control of CRISPR systems greatly improves our ability to achieve highly accurate genomic perturbations spatiotemporally in living cells for cancer treatments.
4.3. Cellular barcoding for cancer therapy
Cellular barcoding involves individual cells being tagged with unique nucleic acid sequences so that they can be tracked through space and time. At present, cellular barcoding has been widely adopted for fate mapping, lineage tracing, and high‐throughput screening, and has greatly contributed to the understanding of developmental biology and gene function. 114 Alejo Rodriguez Fraticelli et al. 115 have traced and characterized the family tree of individual blood cells at the time of their formation in their natural environment by tagging bone marrow cells from mice with barcodes. Their results confirmed that the megakaryocyte lineage was the predominant native fate of long‐term hematopoietic stem cells. Meanwhile, cellular barcoding can also be used for cancer surveillance. Researchers used cellular barcoding of breast cancer xenografts (PDXs) to track the engraftment and growth characteristics of individual cells in early passaged PDX tumors. The authors observed the proliferation characteristics of tumor cells and found that chemotherapy only temporarily reduced the number of harmful cells and did not permanently eliminate them. 116 Such findings significantly improved our understanding of the mechanisms of breast cancer metastasis and drug resistance.
Cellular barcoding technology has been introduced into CRISPR systems to obtain diverse editing patterns and informative loci. Cas9 was known to create insertions or deletions in gene sequences in the absence of a homologous repair template, and these insertions or deletions (genetic “scars”) constituted heritable cellular barcodes that can be read out by scRNA‐seq for genealogical analysis. Spanjaard et al. 117 proposed a strategy that allowed simultaneous lineage tracing and transcriptome analysis in thousands of single cells, known as LINNAEUS (lineage tracing by nuclease‐activated editing of ubiquitous sequences). Injection of Cas9 and sgRNA into single‐cell stage embryos enabled the labeling of genetic scar cells at early developmental stages and allowed for tracing these cells in distinct organ developmental lineages in zebrafish larvae as well as adult fish. Such a strategy provides a new approach to study cancer single‐cell analysis and metastasis. In addition, new strategy with constantly self‐editing DNA barcodes in living cells has been proposed. Cas9 proteins can randomly mutate the sequence of guide RNA in cells, generating homing guide RNA (hgRNA). By performing confusion analysis, the parentage between different cell populations could be determined. When Cas9 was activated, the hgRNA sequence drifted randomly, acting as an evolutionary barcode. Indeed, multiple studies reported that this homing CRISPR–Cas9 system could be used as a genetic barcode, allowing for the controlled sequence diversification, deep lineage tracing, and molecular recording in cancers 118 , 119 (Figure 3c). In recent years, a whole‐organism lineage tracing has been developed, named genome editing of synthetic target arrays for lineage tracing (GESTALT), which can be used to map large‐scale cell lineage in multicellular systems. 120 However, the disadvantages of being limited to early development and the inability to identify cell types restricts its application. 120 Raj et al. 121 , 122 combined single‐cell RNA sequencing of identifiable cell types with GESTALT to generate scGESTALT, which allowed barcodes to be edited at multiple time points. Furthermore, the CRISPR array repair lineage tracing (CARLIN) system was established to simultaneously interrogate the lineage and transcriptomic information of single cells in vivo. Specifically, a universal mouse model was designed by using CARLIN, capable of generating up to 44,000 transcribed barcodes in an inducible manner at any time points during development or adulthood. The CARLIN system is therefore uniquely suited to the study of stem cell cloning dynamics. 123 The abovementioned novel properties make CRISPR barcodes useful for a wide range of applications, such as dissecting cancer biology.
Notably, by establishing MDA‐MB‐231 and AT‐3 cell models with barcoding system and corresponding in vivo mouse models Zhang et al. 124 confirmed the existence of re‐metastasis from bone metastases to other organs during spontaneous metastasis of mammary gland carcinoma in mice. Moreover, multiple combinations of genetic mutations have been implicated in cancer progression and interactions between different mutations are currently unknown. Rogers et al. 125 proposed a parallel approach of tumor barcoding and high‐throughput barcode sequencing (Tuba‐seq) to quantify the impact of many different tumor suppressor gene alterations. By using CRISPR–Cas9‐mediated tumor suppressor inactivation, tumor barcoding, and deep sequencing of DNA barcodes, Tuba‐seq allows us to unravel novel insights into mutant gene interactions in genetically engineered mouse cancer models, and enables us to precisely track every single cell in tumors and significantly accelerate drug development for cancer.
DNA barcoding can label enormous numbers of cells, but can only provide volumetric resolution and does not yield high‐precision phenotypic and somatic cell resolution. 126 To address these issues, a barcoding system that operated at the protein level has been reported. By synthesizing modules encoding linear epitope triplets, more than 100 unique protein barcodes (Pro‐Codes) can be generated, with each Pro‐Code can be paired with a different CRISPR, allowing for simultaneous analysis of multiple cancer phenotypic markers. Employing the Pro‐Code/CRISPR screening system, the authors reported that antigen‐dependent immune editing of cancer cells were controlled by two interferon‐stimulated genes, the immunoproteasome component Psmb8 and a chaperone Rtp4, and identified Socs1 as a negative regulator of Pd‐l1. 127 It is thus clear that Pro‐Code/CRISPR enables phenotypic analysis of a large number of cancer genes with high precision cellular resolution, contributing to the genetic annotation of various cancer. 127 Overall, combining cellular barcoding technology with CRISPR systems has dramatically enriched our understanding of cancers at both the genetic and molecular levels.
4.4. Genome‐wide or CRISPR screens for cancers
Forward genetic screens with genome‐wide CRISPR libraries are powerful tools for resolving cellular circuits and signaling pathways. Also, pooled CRISPR screens are widely used as a method to identify genes involved in biological mechanisms such as cell proliferation, drug resistance, and viral infection. For example, mutations in the MEN1 gene are known to induce tumorigenesis. Using a CRISPR–Cas9 screening system Ma et al. 128 identified dihydroorotic dehydrogenase (DHODH) as a synthetic lethal gene partner of MEN1 in MEN1‐mutated tumor cells. DHODH is primarily found in the inner mitochondrial membrane and is a key enzyme for the de novo synthesis of pyrimidine nucleotides. In malignantly proliferating cells, the cells are dependent on the de novo synthesis pathway to produce sufficient pyrimidine nucleotides. 129 Therefore, inhibition of DHODH in malignant cells can induce cell death. In MEN1 mutation‐associated tumors, DHODH expression was enhanced due to the absence of MEN1 protein, leading to massive proliferation of tumor cells. Excitingly, leflunomide is a potent DHODH inhibitor. 130 By constructing a mouse xenograft model, the researchers administered leflunomide to MEN1 mutation‐associated tumors and showed a significant reduction in tumor size. 128 Meanwhile, through forward genetic screens, CRISPR technologies allow high‐resolution detection of genetic traits in cancer cells. A genome‐wide CRISPR–Cas9 screen of pancreatic ductal adenocarcinoma (PDAC) cells with RNF43 mutations identified the Wnt receptor frizzed‐5 (FZD5) as the only FZD receptor encoded in the human genome, suggesting that antibodies to FZD5 could be used to inhibit PDAC cell proliferation. 131 A genome‐wide CRISPR screen can also be used to reveal the genetic circuits by which cancer cells evade the host immune system. Lawson et al. 132 performed a genome‐wide CRISPR screen on mouse cancer cell lines cultured in the presence of cytotoxic T lymphocytes (CTLs). The results showed that 182 core genes were involved in regulating the sensitivity or resistance of cancer cells to CTL‐mediated toxicity. These studies provide clear evidence to support that CRIPSR‐based genome‐wide genetic screens can be used to identify and validate cell surface targets and core genes for developing antibody treatments and gene therapy for cancers.
Large‐scale screening of complex cellular phenotypes is a shortcoming of the CRISPR screens, and single‐cell RNA sequencing offers a solution to this problem. By combining the advantages of single‐cell RNA sequencing and pooled CRISPR screens, Dixit et al. 133 constructed Perturb‐Seq which allowed the functions of different genes in different cell types in an organism to be studied simultaneously. It has been shown that Perturb‐seq is able to accurately and simultaneously determine the cellular states with distinct gene responses among 200,000 immune cells and in proliferating cell lines. Meanwhile, CRISP‐seq was developed to study the crosstalks and redundancies in complex genetic circuits and massive cellular heterogeneity in multicellular organisms. Specifically, CRISP‐seq has shown clear potential in coding genes, knocking out genes, disrupting genetic elements such as noncoding RNA, promoters, and enhancers, as well as inducing specific gene expression. 134 Moreover, Datlinger et al. 135 developed a CRISPR droplet sequencing method (CROP‐seq) by combining key strengths of pooled and arrayed screens to enable pooled CRISPR screens with single‐cell transcriptome resolution, which will facilitate high‐throughput functional dissection of complex regulatory mechanisms and heterogeneous cell populations (Figure 4a–c). Furthermore, Wheeler et al. 136 reported a method named as CRaft‐ID that combines pooled CRISPR–Cas9 screen with microraft array technology and high‐content imaging to screen image‐based phenotypes. CRaft‐TD extends the application of CRISPR screens to the throughput of image‐based genetic knockout studies, enabling us to study genetic modulators of subcellular and cellular phenotypes.
Moreover, improving the accuracy and robustness of pooled‐library CRISPR screens by capturing sgRNA integrations in individual organoids can significantly reduce required cell numbers for genome‐scale screening. Genetic dissection of the TGF‐β tumor suppressor pathway can be achieved by establishing a genome‐level CRISPR screen for human small intestine (hSI) organoids. 137 Additionally, Michels et al. 138 developed a platform for pooled CRISPR–Cas9 screen in human colon organoids to facilitate high‐throughput genetic testing and functional identification of tumor drivers. In a screened pan‐cancer tumor suppressor gene (TSG) library, TGFBR2 was shown as the most prevalent TSG. The platform for pooled CRISPR–Cas9 screen in human colon organoids permits unbiased detection of genes that confer positive selection in vitro and after xenotransplantation. The successful use of pooled CRISPR screens for human organoids, therefore, provides a powerful platform for identifying patient‐specific vulnerabilities and for genetically dissecting cancer mechanisms in physiologically relevant model systems.
5. CURRENT DELIVERY SYSTEMS OF CRISPR–Cas
The multiple forms and derivatives of the CRISPR–Cas systems as discussed in above sections significantly enrich their applications in cancer at the genetic level. However, the commonly used vector delivery is an urgent issue for CRISPR–Cas systems to enter into human clinical trials. Currently, three main delivery forms are used for delivering CRISPR–Cas9 (Figure 5): (i) Delivering ribonucleoprotein (RNP, a complex of Cas9 protein and sgRNA) 139 , 140 ; (ii) Delivering mRNA and sgRNA components 141 ; (iii) Delivering plasmid DNA encoding Cas9 and sgRNA, such as pX330, pX458, and pX459. 142 , 143 The advantages and disadvantages of these three approaches are shown in Table 2. The CRISPR–Cas systems for targeting DNA must enter the nucleus of the target cell in order to have a therapeutic effect. To achieve such a purpose, a number of delivery systems have been developed for the CRISPR–Cas9 system. Depending on whether viral transduction is used, CRISPR–Cas9 delivery strategies can be broadly classified as viral or nonviral approaches, the latter also includes a variety of physical and chemical delivery strategies (Table 3).
TABLE 2.
Delivery formats | Advantages | Disadvantages | References |
---|---|---|---|
Protein RNP | Rapid; significant reduction in off‐target effects, toxicity, and immune response | Difficult to obtain large amounts of highly active Cas9 protein; large Cas9 protein size | 139 , 140 |
Cas9 mRNA and sgRNA | Faster; favorable for reducing off‐target effects | Poor mRNA stability; poor gene editing | 141 |
Plasmid DNA | Simple; stable; avoiding multiple transfections | Low efficiency; higher off‐target effect; difficult translocation of plasmid DNA to the nucleus | 142 , 143 |
TABLE 3.
Viral delivery | Physical delivery | Chemical delivery | Other emerging nonviral delivery | |||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
Strategy | AAVs | AVs | LVs | Bacteriop‐hages | Microinjection | Electroporation | Hydrodynamic injection | Cationic vectors | CCPs | AuNPs | GO | BP | DNA nanostructure | EVs |
Delivery formats | DNA | DNA | DNA | DNA | DNA, mRNA or protein | DNA, mRNA or protein | DNA, mRNA or protein | DNA, mRNA or protein | Protein | Protein | Protein | Protein | Protein | Protein |
Delivery efficiency | ++ | ++ | +++ | ++ | + | +++ | +++ | + | + | ++ | +++ | +++ | +++ | ++ |
Safety concern | + | ++ | +++ | + | + | + | ++ | + | + | + | + | + | + | + |
Cost | ++ | ++ | + | ++ | +++ | +++ | ++ | + | + | ++ | ++ | ++ | ++ | + |
Technical requirement | ++ | +++ | + | ++ | +++ | + | +++ | + | ++ | ++ | ++ | +++ | +++ | + |
Major applications | in vivo | in vivo | in vitro and ex vivo | in vivo and in vivo | in vitro and ex vivo | in vitro and ex vivo | in vitro and ex vivo | in vitro and in vivo | in vitro and in vivo | in vitro and in vivo | in vitro and in vivo | in vitro and in vivo | in vitro and in vivo | in vitro, ex vivo and in vivo |
Abbreviations: + denotes low; ++ denotes medium; +++ denotes high; AAV, adeno‐associated virus; AuNPs, gold nanoparticles; AV, adenovirus; BP, black phosphorus; CCPs, cell penetrating peptides; EVs, extracellular vesicles; GO, graphene oxide; LV, lentivirus.
5.1. Viral vector delivery
Over 70% of the gene therapy in clinical trials are using viral vector to deliver therapeutic genes, which is accomplished with two main mechanisms: infection and replication. During the infection phase, the virus can identify and enter a specific cell and the viral genome will be released into the nucleus (DNA) or cytoplasm (RNA) for replication. 144 Gene therapy by genome editing is achieved by transporting a virus containing the delivery material (programmed genome editing nuclease) to a target cell. A large number of viral vectors have been developed, mainly adeno‐associated viruses (AAVs), adenoviruses (AVs), lentiviruses (LVs), and bacteriophages. 145 , 146
5.1.1. Adeno‐associated viral vector‐mediated delivery
AAV vectors are a common viral vector used for gene therapy. Compared to other viral vectors, AAV vectors are widely used for delivery of CRISPR–Cas9 system due to their low immunogenicity, long‐lasting transgene expression, mild immune response, high infection efficiency, and general safety. 144 , 147 However, the low packaging capacity of the AAV vectors (<4.7 kb) has limitations for delivery of large size Cas proteins. 148 , 149 To address this limitation, two methods of delivery are now commonly used. The first one is looking for and using smaller‐sized Cas9 proteins. The most commonly used SpCas9 (~4.2 kb) is challenging to deliver, while the smaller derivatives including SaCas9 (~3.2 kb), St1Cas9 (~3.4 kb), and NmeCas9 (~3.2 kb) are feasible for AAV delivery. 150 However, the smaller Cas9 proteins are restricted by the availability of suitable PAM sequences. Second, a dual‐AAV vector system was designed, with one AAV vector delivering spCas9 protein and the other delivering gRNA. 151 Notably, the efficiency of delivery is diminished due to the reduced probability of two AAV vectors delivering to the same cell. Therefore, AAV vectors for delivery of CRISPR–Cas9 system need to be further optimized for greater clinical applicability. In this aspect, Richards et al. 152 engineered a dual‐AAV vector to deliver CRISPR/Cas9 machinery and co‐administered a nonhomologous end‐joining inhibitor vanillin to promote homology‐directed repairs, and found that the Pah enu2 allele was permanently corrected in the hepatocytes of mice with phenylketonuria. A triple AAV carrier has also been engineered to accommodate the larger size of the load and improve the delivery efficiency. 153 Therefore, clinical trials with AAV‐delivered CRISPR/Cas9‐based gene editing are expected to be starting shortly as the AAV delivery system becomes increasingly refined.
5.1.2. Adenovirus‐mediated delivery
AVs are a nonenveloped double‐stranded DNA virus that could transduce both dividing and nondividing cells. In addition, they have no endogenous integration mechanism and do not integrate into the host cell genome. 154 AVs are widely used in clinical trials for gene therapy because of their good biological properties, genetic stability, and high gene transfer efficiency. 155 Kato et al. 156 developed an AV vector that could edit the X gene in HBV cell lines and showed that the AV vector successfully blocked the HBV‐X gene in heterogeneous patients. However, AVs tend to trigger innate immune responses within host cells, leading to tissue inflammation. This, coupled with the laborious production of AVs, further limits the efficiency of clinical application. 157 , 158
5.1.3. Lentiviral vector‐mediated delivery
LVs derived from HIV‐1 viruses have tropism and larger cargo capacity (~8 kb), which allow Cas9 and gRNA to be loaded into an LV vector simultaneously. 159 LVs have the ability to transduce nondividing cells efficiently compared to other viral vectors, making them an instrumental tool for gene therapies in somatic and germline cells. 160 Since all viral genes are deleted, the LV vector does not activate the immune system. 161 Furthermore, the production of LVs is simpler than that of AAVs and AVs. 162 Taking these advantages, Holmgaard et al. 163 successfully delivered SpCas9 protein and sgRNA via LV vector to achieve Vegfa genome editing in the mouse retina. Similar to AVV vectors, dual‐LV vector systems were developed for efficient, time‐controlled deletion of MCL‐1 (a gene controls apoptosis, making tumor cells resistant to chemotherapeutic agents) in vitro and in vivo. 164 However, as a retrovirus, the host genome integration of LVs leads to unwanted off‐target insertional mutagenesis that may result in the inactivation of tumor suppressor genes or the activation of proto‐oncogenes and thus induce tumorigenesis. 165 , 166 Clinical trials of LV vector‐delivered CRISPR–Cas systems are being shelved due to potential risks of tumorigenesis.
5.1.4. Bacteriophage‐mediated delivery
Bacteriophages are a group of viruses that infect bacteria or archaea and are often used against multidrug‐resistant bacteria. However, the specific susceptivity of bacteria has limited the development of bacteriophage therapy. Fortunately, the combination of bacteriophages and the CRISPR–Cas system has changed this dilemma. Bikard et al. 167 designed a plasmid containing the cas9 gene and its gRNA sequence, and delivered it to S. aureus using bacteriophage. The results showed that the delivery system was effective in killing S. aureus on the skin surface of mice. Subsequently, various CRISPR–Cas systems using bacteriophage delivery have been widely used for controlling bacterial infections. Selle et al. 168 used bacteriophages to successfully deliver CRISPR–Cas3 and significantly inhibit C. difficile infection. Using a similar strategy, Liu et al. 169 delivered the bacteriophage‐mediated CRISPR/Cas9 to drug‐resistant bacteria and showed a sustained and efficient eradication of drug‐resistant bacteria. Despite the huge success in controlling bacterial infection, there is little evidence to show the application of bacteriophage‐mediated CRISPR–Cas delivery and gene editing in cancer treatment.
5.2. Nonviral delivery
Several nonviral vectors have been developed and successfully applied for CRISPR–Cas delivery (either DNA, mRNA, or protein), such as physical methods to disrupt the cellular barriers, chemical modifications to improve cargo transport to avoid the barriers, and physical encapsulation of the cargo in the vector molecule. 170 The main advantages of nonviral vectors are their ability to accommodate large size delivery, ease of generation, good controllability, and safety.
5.2.1. Physical delivery
Delivery of CRISPR–Cas is achieved by exposing cells to mild physical conditions that temporarily disrupt the physical barriers that prevent the cargo from reaching its intended destination. Microinjection is a physical method of injecting Cas9 and sgRNAs directly into cells using a microscope and a microinjection needle. 171 This method has been used for DNA, RNA, or RNP delivery of Cas9, as well as for direct delivery of gRNAs. 171 , 172 , 173 However, the method is heavily relied on well‐established experimental facilities and delicate handling to avoid permanent damage to the membrane, and is therefore commonly used for in vitro cell experiments. Electroporation is the temporary disturbance of the lipid bilayer of the plasma membrane by an electric field, thereby enhancing the permeability of the cell membrane. 174 The method has been successful in delivering DNA, RNA, and even RNPs in vitro. 139 , 175 , 176 However, electroporation‐mediated gene editing is costly and requires specific induction conditions to be set for different cell types. Importantly, high level of cell death caused by electroporation also limits its clinical application. 177 Both physical methods have been widely used in vitro, but only a few strategies have been used for in vivo delivery of the CRISPR–Cas system, including hydrodynamic injection. 178 Hydrodynamic injection enables delivery of CRISPR–Cas by creating temporary pores in the cell membrane at high pressure for a short period of time. 179 For example, using hydrodynamic injection of pX330‐PTEN plasmid, Yu et al. 180 effectively decreased the expression level of PTEN in rat liver and successfully induced lipid deposition and nonalcoholic fatty liver. Currently, the method can only be applied to small animal models due to the large starting injection volume. 179 Therefore, it is not currently suitable for human application.
A number of new physical methods are being developed for CRISPR–Cas component delivery. Sessions et al. 181 proposed a scheme, named lance array nanoinjection (LAN) that produced transient pores (1–2.5 μm diameter) in the target cell membrane. The results showed successful transfer of the CRISPR/Cas9 plasmid into the GFP+/FRT HeLa cell line and the eventual modification of the EGFP gene. Importantly, LAN did not cause massive cell death, unlike with electroporation. Additionally, ultrasonically driven nanomotors have exhibited direct and rapid cell membrane permeation and have been applied to the intracellular delivery of ncRNA. 182 Wang et al. 183 loaded Cas9/sgRNA complex onto the surface of nanomotors by disulfide bond modification. The obtained nanomotors could directly penetrate the cytoplasm of GFP‐expressing mouse melanoma B16F10 cells after ultrasound treatment. The results showed that the GFP gene was efficiently knocked out. Taken together, these newly developed physical methods offer potential alternative strategies for delivery of the CRISPR–Cas systems to tumor cells.
5.2.2. Chemical delivery
Physical methods are readily available in the laboratory setting, but less practicable for clinical application. Chemical deliveries use complementary components to assist Cas9 in bypassing cellular barriers and protect it from degradation. There are two main forms of chemical delivery: encapsulation of cargos in another chemical entity, or direct chemical modification of cargos, which mainly includes cationic vectors [such as lipid nanoparticles (LNP), cell penetrating peptides (CPPs), and inorganic metal nanoparticles (NPs)]. 184 , 185 , 186
As both nucleic acids and cell membranes are negatively charged, the repulsive forces between them prevent nucleic acids from entering the cell. Cationic modifications can, on the one hand, mediate binding to the nucleic acid through electrostatic interactions and protect the nucleic acid from destruction by nucleases. On the other hand, it assists the nucleic acid to cross the cell membrane and complete the delivery of the cargo. 187 , 188 Currently, there are two main categories of cationic‐based vectors: cationic lipid‐based vectors (LNP) and cationic polymer‐based vectors. LNP, in which negatively charged nucleic acids can be encapsulated, has been widely used for the delivery of ncRNAs and the like. This method is less efficient in delivery, but is safe, simple, and inexpensive. 31 LNP has also been used to deliver CRISPR–Cas9 system. For example, Zuris et al. 189 successfully transfected gRNA and Cas9 RNP complexes into a HeLa reporter cell line via cationic liposomes, achieving an 80% gene modification rate. Moreover, sgRNA targeting the DPP‐4 gene was designed to form a RNP with Cas9 protein and encapsulated with lecithin cationic liposomes. The results showed that the nano‐preparation resulted in the destruction of the DPP‐4 gene in type 2 diabetic mice. 190 Recently, encapsulation of focal adhesion kinase (FAK) siRNA, Cas9 mRNA, and sgRNA (siFAK + CRISPR‐LNPs) in LNP achieved a 10‐fold enhancement of gene editing in tumor spheroids. By further constructing siFAK + CRISPR‐PD‐L1‐LNPs, the expression of PD‐L1 was effectively blocked, and the tumor growth and metastasis of the four cancer mouse models were also significantly inhibited. 191 LNP can also be combined with physical methods. Yin et al. 192 combined the LNP delivering Cas9 mRNA with AAVs encoding a sgRNA and a repair template to form a recombinant template after hydrodynamic tail vein injection, increasing repair efficiency from 0.4% to 6%. However, these vectors are currently mainly used for gene editing at the cellular level since they do not contain reactive functional groups, with poor tissue and cellular targeting ability and low in vivo stability.
Cationic polymer‐based vectors offer a wider range of options and more flexible structural design than cationic lipid vectors. Polyethyleneimine (PEI) is a commonly used polymer for CRISPR–Cas9 delivery. Ryu et al. 193 found that branched PEI 25 kDa (BPEI‐25K) was effective in delivering plasmids expressing gRNA and Cas9 into Neuro2a cells for successful gene editing of Slc26a4 locus. Furthermore, the joint design of multiple polymers further expands the tool box. Luo et al. 194 encapsulated Cas9 expression plasmids (pM458 and pM330) in cationic lipid‐assisted PEG‐b‐PLGA nanoparticles (CLAN) to achieve gene editing in macrophages in vivo and in vitro, respectively. By designing sgRNAs targeting Ntn1 gene, CLAN successfully disrupted the Ntn1 gene in macrophages and their precursor monocytes in vivo, ameliorating the symptoms of T2D. However, the issue of cytotoxicity still needs further optimization.
Cell penetrating peptides (CPPs) are short peptides with the ability to across cell membranes. CPPs are suitable for preclinical and clinical studies due to their low cytotoxicity compared to other vectors and their eventual degradation to amino acids. 185 They are currently used as tools to achieve efficient Cas9 protein and sgRNA delivery. 195 , 196 Ramakrishna et al. 197 conjugated Cas9 protein to CPP via a thioether bond, whereas gRNA bound with CPP to form positively charged NPs. The NPs resulted in efficient gene disruption and low off‐target mutation rates compared to plasmid transfections in embryonic stem cells, dermal fibroblasts, HEK293T cells, HeLa cells, and embryonic carcinoma cells. Moreover, Suresh et al. 196 delivered the CRISPR–Cas9 system using CPP‐conjugated recombinant Cas9 protein and CPP‐complexed gRNAs to disrupt gene expression in human cell lines. Meanwhile, an amphiphilic penetrating peptide formed via a hydrazone bond formation between a cationic peptide scaffold and a hydrophobic aldehyde tail was designed for direct delivery of Cas9, with results showing that 30%–40% gene editing efficiency was obtained. 198 These results have collectively indicated that CPPs have great potential for applications in CRISPR–Cas9 system. However, there are some limitations for CPPs, including a short half‐life in the blood and nonspecific delivery. 199 Therefore, improving the specificity and efficiency of CPPs is the direction for future breakthroughs, especially in tumor cells.
Inorganic nanomaterials can also be used for vector delivery. Gold nanoparticles (AuNPs) are considered as good vectors for gene delivery as they are easy to control in terms of size and distribution, have good chemical stability and biosafety, and can be chemically modified. 200 , 201 Mout et al. 202 designed cationic arginine gold nanoparticles (ArgNPs) that successfully delivered engineered Cas9 protein and sgRNA to HEK‐293T and Raw 264.7 cell lines, and achieved a 90% delivery efficiency and 30% gene editing efficiency. Furthermore, a light‐controlled release nano‐delivery system (LACP) based on AuNPs‐plasmids was developed for the delivery of Cas9‐sgPLK‐1 plasmids (CP). The results showed effective knockdown of the PLK‐1 gene in melanoma and inhibition of tumor growth in vitro and in vivo. 203 Tao et al. 204 were the first to report the application of gold nanoclusters (AuNCs) encapsulated with protamine as Cas9‐sgRNA plasmid vectors, and found that AuNCs were able to rapidly assemble with Cas9‐sgRNA plasmids, while the cationic protamine facilitated the efficient release of Cas9‐sgRNA plasmids in the nucleus. Therefore, AuNPs provide a safe delivery method for CRISPR–Cas components and are expected to play a greater role in the clinical treatment of tumors.
5.3. Other emerging nonviral delivery
Graphene oxide (GO) and black phosphorus (BP) nanosheets can also be used for CRISPR–Cas9 delivery. GO has good bio‐compatibility and safety, and is capable of loading Cas9/sgRNA RNP after modification with PEG and PEI. GO‐PEG‐PEI can rapidly transfer RNP into human cells while retaining Cas9 activity, with a gene editing efficiency of 39%. 205 Besides, BPs, also known as isotopes of elemental phosphorus, have good biocompatibility. Zhou et al. 206 loaded Cas9 RNP onto BP nanosheets by electrostatic adsorption. Under acidic environment, BP can be rapidly degraded to biocompatible inorganic phosphate and RNP is released into the nucleus, ultimately achieving efficient gene editing and gene silencing.
In recent years, the intrinsic properties between DNAs allow DNA nanostructures to self‐assemble in precisely controlled sizes and shapes. Currently, DNA nanostructures are considered as an emerging delivery system with the advantages of high payload capacity, good biocompatibility and biodegradability, and high stability under physiological conditions. 207 , 208 , 209 Sun et al. 210 synthesized yarn‐like DNA nanoparticles that can efficiently load Cas9/sgRNA complexes, and successfully delivered to the nucleus of human cells to achieve EGFP gene knockdown. DNA nanoparticles have not been widely used for delivery of CRISPR–Cas components, probably due to the complexity of the operation.
Apart from the abovementioned novel nonviral delivery systems, extracellular vesicles (EVs) have been shown to carry a variety of biomolecules (plasmids, siRNAs, and miRNAs) due to their high biocompatibility and low immunogenicity. 211 In contrast to lentiviral delivery systems, EVs do not contain any viral genome. Therefore, EVs do not integrate into the host genome. 212 In addition, EVs release Cas9 transiently in cells, greatly reducing the chance of off‐target effects due to long‐term Cas9 expression, 213 making them the ideal system for the delivery of CRISPR–Cas9 components. Indeed, two main approaches have been reported to improve the encapsulation of CRISPR–Cas9 by EVs: engineering EVs or introducing CRISPR–Cas9‐enriched components that efficiently load mRNA and RNP into EVs through molecular interactions, 214 , 215 , 216 confirming the applicability of EVs for in vivo gene therapy.
6. CLINICAL APPLICATION AND CHALLENGES IN CANCER
Today, CRISPR technologies have been applied to a range of scientific studies such as precision genome editing and transcriptional regulation, enabling scientists to manipulate gene sequences “at will”. In addition, CRISPR‐based gene therapies are already being used in clinical trials, particularly for diseases caused by genetic mutations. 217 , 218 One of the most effective ways to study the function of these mutations is to create models carrying mutated genes. Importantly, with the help of CRISPR technology, a large number of cancer models can be generated by modifying key cancer‐related genes. By developing cell‐specific viral and nonviral vectors, specific cell can be precisely manipulated, and thus cellular and animal models can be constructed. 219 For example, several genes could be successfully modified in single mouse hematopoietic stem cells to recapitulate genetic complexity in human malignancies by delivering sgRNAs and Cas9 with a lentiviral vector, as exemplified by the creation of acute myeloid leukemia murine model through simultaneously mutating five genes in a single mouse hematopoietic stem cell. 220 Similarly, PTEN and p53 are known as two tumor suppressor genes. 221 , 222 Accordingly, a murine model of liver tumors was successfully constructed by Xue et al. 223 through delivering a CRISPR plasmid DNA expressing Cas9 and sgRNAs to liver by hydrodynamic injection to directly target and inactivate PTEN and p53 gene expressions. Moreover, Maresch et al. 224 successfully induced pancreatic cancer models by delivering CRISPR–Cas9 vectors directly into the pancreas of mice. These studies have collectively demonstrated that CRISPR systems mediated gene editing can be efficiently used for establishing a variety of cancer models, which greatly facilitates study of molecular mechanisms of cancer and therapeutics for cancer therapy.
CRISPR systems have shown great potential not only for the construction of cancer models but also for the treatment of cancer, especially in cancer immunotherapy. When cancer cells invade, immune cells can remove them through “immune surveillance”. However, cancer cells have evolved an “immune escape” mechanism to survive. Cancer immunotherapy (also called “immuno‐oncology”) is used to enhance the anti‐cancer ability of immune cells so that cancer cells cannot complete immune escape. 225 Anti‐programmed death 1 (PD‐1) antibodies are one of the most studied and fastest growing immunotherapies in clinical practice. In 2016, Chinese scientists conducted the world's first human CRISPR clinical trial (NCT02793856) to confirm the safety and feasibility of clinical application of CRISPR–Cas9. Specifically, researchers extracted T cells from nonsmall cell lung cancer patients and co‐transfected Cas9 and sgRNA plasmids into isolated T cells by electroporation, thereby editing the PD‐1 gene of the T cells. The amplified engineered cells were then reinfused into the patients. The results showed a median progression‐free survival of 7.7 weeks (95% confidence interval, 6.9–8.5 weeks) and a median overall survival was 42.6 weeks (95% confidence interval, 10.3–74.9 weeks) in 12 patients. In addition, the median mutation frequency of off‐target events was 0.05% (range 0%–0.25%). 218
However, immunotherapies that block PD‐1 checkpoints have sometimes shown nonresponsiveness. Therefore, CRISPR is often used to implement cancer immunotherapy by combining synthetic biology to screen the genome. Manguso et al. 226 screened 2368 genes in melanoma using CRISPR and identified PTPN2 as a new immunotherapy target. Upon PTPN2 knockdown, IFN‐γ‐mediated antigen presentation and growth inhibition were enhanced, further improving the efficacy of immunotherapy. Similarly, in animal models of triple‐negative breast cancer (TNBC), the immune efficacy of CD8 T cells against TNBC was enhanced following the identification of Dhx37 knockdown by CRISPR screening of CD8 T cells. 227 Using the same approach, Wang et al. 228 identified in TNBC mouse models that deletion of the E3 ubiquitin ligase Cop1 resulted in reduced tumor macrophage infiltration and enhanced antitumor immunity.
As the application of immune checkpoint inhibitors gradually expands, new advances in CAR‐T therapy research continue to emerge. A similar clinical trial using CRISPR gene editing T‐cell was also conducted in the U.S (NCT03399448). The researchers used CRISPR–Cas9 to knock out three genes (TRAC, TRBC, and PDCD1) from the patients' T cells, and such modified T cells (CAR‐T cells) were reinfused into three patients with refractory cancer. The results showed that the edited T cells were survived in vivo for up to 9 months without any serious adverse events, demonstrating the feasibility of CRISPR–Cas9 for cancer immunotherapy. 229
Although autologous CAR‐T cells have shown promising results in cancer treatment, there are still some barriers to overcome, particularly the extreme high cost for personalized VAR‐T cell production. Excitingly, universal CAR‐T cells from healthy donors offer a strategy for large‐scale clinical application. In an open‐label dose‐escalation phase I study, Hu et al. 230 developed universal CD19/CD22‐targeting CAR‐T cells (CTA101) using CRISPR/Cas9 to disrupt TRAC region and CD52 gene. After 28 days, the complete remission (CR) rate in six patients who received CTA101 infusion was 83.3%. Importantly, PD‐1 blockade could potentially enhance the immunotherapeutic efficacy of CAR T cells. It was verified that CAR T cells were able to recognize mesothelin overexpressed in human TNBC cells. Meanwhile, after knocking down the PD‐1 motif in human primary T cells using CRISPR/Cas9, CAR T cell cytokine production and cytotoxicity against PD‐L1‐expressing cancer cells were strongly enhanced. In vivo experimental results further revealed that tumor growth was significantly inhibited. 231 Thus, combined immune checkpoint blockade with CAR T cells may enhance cancer immunotherapy and provide a strategy for controlling solid tumor progression.
Currently, three CRISPR–Cas gene editing based allogeneic CAR‐T cell therapies are in cancer clinical trials (NCT04035434, NCT03398967, and NCT03166878). In addition, enabling in vivo editing can further expand the applicability of cancer treatment. The team from the First Affiliated Hospital of Sun Yat‐Sen University piggybacked the CRIPR‐Cas9 plasmid into a gel containing C32‐447 and Poloxamer 407, which in turn disrupted the E6/E7 DNA of HPV16 and HPV18 (NCT03057912). The efficacy of in vivo editing is influenced by the precision of delivery. Fortunately, delivery strategies for CRISPR–Cas systems are being refined at an accelerated pace. Currently, CRISPR–Cas systems are already available for clinical trials of various types of cancer, and the outcome will be emerging shortly (Table 4). Collectively, the CRISPR–Cas system represents a powerful and promising tool in cancer modeling, diagnosis and treatment.
TABLE 4.
Target gene | Condition or disease | Phase | Status | First Posted | ClinicalTrials.gov Identifier | Sponsor |
---|---|---|---|---|---|---|
PD‐1 and TCR | Solid tumor, adult | Phase I | Recruiting | June 4, 2018 | NCT03545815 | Chinese PLA General Hospital |
CISH | Gastrointestinal cancers | Phase I/II | Recruiting | June 11, 2020 | NCT04426669 | Intima Bioscience, Inc. |
E6 and E7 | HPV‐related malignant neoplasm | Phase I | Unknown | February 20, 2017 | NCT03057912 | First Affiliated Hospital, Sun Yat‐Sen University |
PD‐1 | Solid tumor, adult | Phase I | Unknown | November 20, 2018 | NCT03747965 | Chinese PLA General Hospital |
CD19 |
B‐cell malignancy; Non‐Hodgkin lymphoma; B‐cell lymphoma |
Phase I | Recruiting | July 29, 2019 | NCT04035434 | CRISPR Therapeutics AG |
PD‐1 | Esophageal cancer | Not Applicable | Completed | March 16, 2017 | NCT03081715 | Hangzhou Cancer Hospital |
PD‐1 | Metastatic nonsmall cell lung cancer | Phase I | Completed | June 8, 2016 | NCT02793856 | Sichuan University |
NF1 |
Neurofibromatosis type 1; Tumors of the central nervous system |
Not Applicable | Suspended | November 6, 2017 | NCT03332030 | Roger Packer |
CD70 | T‐cell lymphoma | Phase I | Recruiting | August 6, 2020 | NCT04502446 | CRISPR Therapeutics AG |
TP53 | High grade ovarian serous adenocarcinoma | Not Applicable | Recruiting | July 30, 2018 | NCT03606486 | University of Washington |
TGF‐β receptor II | Solid tumor, adult | Phase I | Not yet recruiting | July 26, 2021 | NCT04976218 | Chinese PLA General Hospital |
CD19 |
Leukemia lymphocytic acute (ALL) in relapse; Leukemia lymphocytic acute (All) Refractory; Lymphoma, B‐cell |
Phase I | Recruiting | July 30, 2019 | NCT04037566 | Xijing Hospital |
CD5 |
CD5+ relapsed/refractory hematopoietic malignancies; Chronic lymphocytic leukemia (CLL); Mantle cell lymphoma (MCL); Diffuse large B‐cell lymphoma (DLBCL); Follicular lymphoma (FL); Peripheral T‐cell lymphomas (PTCL) |
Early Phase I | Not yet recruiting | February 23, 2021 | NCT04767308 | Huazhong University of Science and Technology |
BCMA | Multiple myeloma | Phase I | Recruiting | January 28, 2020 | NCT04244656 | CRISPR Therapeutics AG |
PD‐1 |
Stage IV gastric carcinoma; Stage IV nasopharyngeal carcinoma; T‐cell lymphoma stage IV; Stage IV adult Hodgkin lymphoma; Stage IV diffuse large B‐cell lymphoma |
Phase I/II | Recruiting | February 7, 2017 | NCT03044743 | Yang Yang |
NTLA‐5001 | Acute myeloid leukemia | Phase I/II | Recruiting | October 4, 2021 | NCT05066165 | Intellia Therapeutics |
TCR and B2M |
B‐cell leukemia; B‐cell lymphoma |
Phase I/II | Recruiting | May 25, 2017 | NCT03166878 | Chinese PLA General Hospital |
CD70 | Renal cell carcinoma | Phase I | Recruiting | June 18, 2020 | NCT04438083 | CRISPR Therapeutics AG |
CD19 |
Non‐Hodgkin lymphoma; Lymphoma; B‐cell lymphoma; B‐cell non‐Hodgkin's lymphoma |
Phase I | Recruiting | November 20, 2020 | NCT04637763 | Caribou Biosciences, Inc. |
CD19 |
Acute lymphoblastic leukemia; Chronic lymphocytic leukemia; Non‐Hodgkin lymphoma |
Phase I | Not yet recruiting | September 8, 2021 | NCT05037669 | University of Pennsylvania |
CD52 and TRAC | B acute lymphoblastic leukemia | Phase I | Recruiting | September 21, 2020 | NCT04557436 | Great Ormond Street Hospital for Children NHS Foundation Trust |
CD19, CD20 and CD22 |
B‐cell leukemia; B‐cell lymphoma |
Phase I/II | Recruiting | January 16, 2018 | NCT03398967 | Chinese PLA General Hospital |
PD‐1 | Advanced hepatocellular carcinoma | Phase I | Recruiting | June 5, 2020 | NCT04417764 | Central South University |
Note: The information is from ClinicalTrials.gov, accessed in February 2022.
However, there are still some challenges associated with CRISPR technologies remain to be circumvented before moving to clinical applications, including adaptability, editing efficiency, delivery methods, off‐target effects, and potential on‐target mutagenesis. 232 , 233 For example, the p53 gene is known as the “guardian of the genome”. DNA double‐strand breaks could be recognized by the p53 gene, which in turn prevents cell division and corrects the error, thus affecting editing efficiency of CRISPR systems. It has been shown that CRISPR–Cas9 gene editing can activate the p53‐mediated DNA damage response, which increases potential risks. 234 These issues must therefore be addressed if CRISPR systems are to be used to precisely target cancer‐related genes in human patients. Excitingly, the advent of extended toolkits and synthetic biology offers a solution to these issues. By artificially controlling various components, the behavior of molecules can be precisely regulated, allowing us to efficiently circumvent the shortcomings of CRISPR systems during cancer diagnosis and treatment. For instance, Huang et al. 235 applied logic circuits and optogenetic devices to the split‐dCas9 system in an attempt to inhibit bladder cancer progression. The system used AND logic gates with the hTERT and hUP II promoters as activation keys. Using a similar strategy, the expression of p53 or E‐cadherin protein could be activated by induction of blue light, thereby inhibiting tumor cell proliferation. Altogether, the combined use of one or more synthetic biology strategies offers a potential therapeutic intervention for cancer treatments.
7. CONCLUSIONS
The CRISPR–Cas system is considered as a powerful tool for cancer treatment due to its robust gene editing capabilities. To date, various types of CRISPR–Cas systems have been developed to be suitable for treating different cancer types. Multifunctional modifications of Cas effector proteins have enhanced the targeting specificity and expanded the targeting range of CRISPR–Cas systems. Incorporating synthetic biology into the CRISPR–Cas system allows us to engineer the CRISPR–Cas systems, making them suitable for clinical application in cancer treatment. Moreover, various newly established delivery strategies made it possible to achieve specific genome editing in vivo, paving the way for successful genetic corrections in human via in vivo gene editing. A number of CRISPR‐based therapies are currently in clinical trials, and we are eagerly waiting for the outcomes to be reported shortly. It is important to note that majority of the endpoints for current ongoing clinical trials are mainly focused on safety and feasibility of CRISPR‐based therapies. More preclinical and clinical data are needed with a view to the future use of CRISPR technology for clinical treatment of a wide range of human cancers.
AUTHOR CONTRIBUTIONS
Xiang Meng: Data curation (lead); investigation (lead); methodology (lead); visualization (lead); writing – original draft (lead). Tian‐gang Wu: Data curation (equal); investigation (equal); methodology (lead); visualization (lead); writing – original draft (equal). Qiu‐yue Lou: Data curation (equal); investigation (equal); methodology (equal); visualization (equal); writing – original draft (equal). Kai‐yuan Niu: Formal analysis (equal); investigation (equal); software (equal). Lei Jiang: Investigation (equal); software (equal). Qing‐zhong Xiao: Conceptualization (equal); project administration (equal); resources (equal); supervision (equal); writing – review and editing (equal). Tao Xu: Conceptualization (equal); funding acquisition (equal); project administration (equal); resources (equal); supervision (equal). Lei Zhang: Conceptualization (equal); funding acquisition (equal); project administration (equal); resources (equal); supervision (equal); writing – review and editing (equal).
FUNDING INFORMATION
The present work was supported by the Natural Science Foundation of China [81700522]; the University Natural Science Research Project of Anhui Province [KJ2021A0272]; 2021 Disciplinary Construction Project in the School of Dentistry, Anhui Medical University [2021kqxkFY06]; the School Research Fund Project of Anhui Medical University [2021xkj127].
CONFLICT OF INTEREST
The authors declare that they have no conflict of interest.
PEER REVIEW
The peer review history for this article is available at https://publons.com/publon/10.1002/btm2.10474.
Meng X, Wu T, Lou Q, et al. Optimization of CRISPR–Cas system for clinical cancer therapy. Bioeng Transl Med. 2023;8(2):e10474. doi: 10.1002/btm2.10474
Xiang Menga, Tian‐gang Wua, and Qiu‐yue Lou contributed equally to this work.
Funding information 2021 Disciplinary Construction Project in School of Dentistry, Anhui Medical University, Grant/Award Number: 2021kqxkFY06; National Natural Science Foundation of China, Grant/Award Number: 81700522; University Natural Science Research Project of Anhui Province, Grant/Award Number: KJ2021A0272
Contributor Information
Qing‐zhong Xiao, Email: q.xiao@qmul.ac.uk.
Tao Xu, Email: xutao@ahmu.edu.cn.
Lei Zhang, Email: zhanglei6551@126.com.
DATA AVAILABILITY STATEMENT
The data that support the findings of this study are available from the corresponding author upon reasonable request.
REFERENCES
- 1. Sung H, Ferlay J, Siegel RL, et al. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2021;71:209‐249. [DOI] [PubMed] [Google Scholar]
- 2. Bu W, Wang Z, Meng L, et al. Disulfiram inhibits epithelial‐mesenchymal transition through TGFβ‐ERK‐snail pathway independently of Smad4 to decrease oral squamous cell carcinoma metastasis. Cancer Manag Res. 2019;11:3887‐3898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Meng X, Lou QY, Yang WY, et al. The role of non‐coding RNAs in drug resistance of oral squamous cell carcinoma and therapeutic potential. Cancer Commun. 2021;41:981‐1006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Hudson TJ, Anderson W, Artez A, et al. International network of cancer genome projects. Nature. 2010;464:993‐998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Gaj T, Gersbach CA, Barbas CF 3rd. ZFN, TALEN, and CRISPR/Cas‐based methods for genome engineering. Trends Biotechnol. 2013;31:397‐405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Zhang HX, Zhang Y, Yin H. Genome editing with mRNA encoding ZFN, TALEN, and Cas9. Mol Ther. 2019;27:735‐746. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Waryah CB, Moses C, Arooj M, et al. Zinc fingers, TALEs, and CRISPR systems: a comparison of tools for epigenome editing. Methods Mol Biol. 2018;1767:19‐63. [DOI] [PubMed] [Google Scholar]
- 8. Mollanoori H, Teimourian S. Therapeutic applications of CRISPR/Cas9 system in gene therapy. Biotechnol Lett. 2018;40:907‐914. [DOI] [PubMed] [Google Scholar]
- 9. Yan J, Jia Y, Chen H, et al. Long non‐coding RNA PXN‐AS1 suppresses pancreatic cancer progression by acting as a competing endogenous RNA of miR‐3064 to upregulate PIP4K2B expression. J Exp Clin Cancer Res. 2019;38:390. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Katti A, Diaz BJ, Caragine CM, et al. CRISPR in cancer biology and therapy. Nat Rev Cancer. 2022;22:259‐279. [DOI] [PubMed] [Google Scholar]
- 11. Marraffini LA. CRISPR‐Cas immunity in prokaryotes. Nature. 2015;526:55‐61. [DOI] [PubMed] [Google Scholar]
- 12. Wright AV, Nuñez JK, Doudna JA. Biology and applications of CRISPR systems: harnessing Nature's toolbox for genome engineering. Cell. 2016;164:29‐44. [DOI] [PubMed] [Google Scholar]
- 13. Sorek R, Kunin V, Hugenholtz P. CRISPR‐‐a widespread system that provides acquired resistance against phages in bacteria and archaea. Nat Rev Microbiol. 2008;6:181‐186. [DOI] [PubMed] [Google Scholar]
- 14. Bhaya D, Davison M, Barrangou R. CRISPR‐Cas systems in bacteria and archaea: versatile small RNAs for adaptive defense and regulation. Annu Rev Genet. 2011;45:273‐297. [DOI] [PubMed] [Google Scholar]
- 15. Cui YR, Wang SJ, Chen J, et al. Allosteric inhibition of CRISPR‐Cas9 by bacteriophage‐derived peptides. Genome Biol. 2020;21:51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Görücü YS. Genome editing technologies: CRISPR, LEAPER, RESTORE, ARCUT, SATI, and RESCUE. EXCLI J. 2021;20:19‐45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Shmakov S, Smargon A, Scott D, et al. Diversity and evolution of class 2 CRISPR‐Cas systems. Nat Rev Microbiol. 2017;15:169‐182. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Koonin EV, Makarova KS, Zhang F. Diversity, classification and evolution of CRISPR‐Cas systems. Curr Opin Microbiol. 2017;37:67‐78. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Makarova KS, Wolf YI, Iranzo J, et al. Evolutionary classification of CRISPR‐Cas systems: a burst of class 2 and derived variants. Nat Rev Microbiol. 2020;18:67‐83. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Shmakov S, Abudayyeh OO, Makarova KS, et al. Discovery and functional characterization of diverse class 2 CRISPR‐Cas systems. Mol Cell. 2015;60:385‐397. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Pickar‐Oliver A, Black JB, Lewis MM, et al. Targeted transcriptional modulation with type I CRISPR‐Cas systems in human cells. Nat Biotechnol. 2019;37:1493‐1501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Nuñez JK, Kranzusch PJ, Noeske J, et al. Cas1‐Cas2 complex formation mediates spacer acquisition during CRISPR‐Cas adaptive immunity. Nat Struct Mol Biol. 2014;21:528‐534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Levy A, Goren MG, Yosef I, et al. CRISPR adaptation biases explain preference for acquisition of foreign DNA. Nature. 2015;520:505‐510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Shin U, Brondani V. Analysis of wild type LbCpf1 protein, and PAM recognition variants, in a cellular context. Front Genet. 2020;11:571591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Amitai G, Sorek R. CRISPR‐Cas adaptation: insights into the mechanism of action. Nat Rev Microbiol. 2016;14:67‐76. [DOI] [PubMed] [Google Scholar]
- 26. Jackson SA, McKenzie RE, Fagerlund RD, et al. CRISPR‐Cas: adapting to change. Science. 2017;356:eaal5056. [DOI] [PubMed] [Google Scholar]
- 27. Chylinski K, Le Rhun A, Charpentier E. The tracrRNA and Cas9 families of type II CRISPR‐Cas immunity systems. RNA Biol. 2013;10:726‐737. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Hille F, Richter H, Wong SP, et al. The biology of CRISPR‐Cas: backward and forward. Cell. 2018;172:1239‐1259. [DOI] [PubMed] [Google Scholar]
- 29. Marraffini LA, Sontheimer EJ. CRISPR interference: RNA‐directed adaptive immunity in bacteria and archaea. Nat Rev Genet. 2010;11:181‐190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Chen H, Choi J, Bailey S. Cut site selection by the two nuclease domains of the Cas9 RNA‐guided endonuclease. J Biol Chem. 2014;289:13284‐13294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Karimian A, Azizian K, Parsian H, et al. CRISPR/Cas9 technology as a potent molecular tool for gene therapy. J Cell Physiol. 2019;234:12267‐12277. [DOI] [PubMed] [Google Scholar]
- 32. Jinek M, Chylinski K, Fonfara I, et al. A programmable dual‐RNA‐guided DNA endonuclease in adaptive bacterial immunity. Science. 2012;337:816‐821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Hendriks D, Clevers H, Artegiani B. CRISPR‐Cas tools and their application in genetic engineering of human stem cells and organoids. Cell Stem Cell. 2020;27:705‐731. [DOI] [PubMed] [Google Scholar]
- 34. Fedorova I, Arseniev A, Selkova P, et al. DNA targeting by clostridium cellulolyticum CRISPR‐Cas9 type II‐C system. Nucleic Acids Res. 2020;48:2026‐2034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Bubeck F, Hoffmann MD, Harteveld Z, et al. Engineered anti‐CRISPR proteins for optogenetic control of CRISPR‐Cas9. Nat Methods. 2018;15:924‐927. [DOI] [PubMed] [Google Scholar]
- 36. Lin P, Qin S, Pu Q, et al. CRISPR‐Cas13 inhibitors block RNA editing in bacteria and mammalian cells. Mol Cell. 2020;78:850‐61.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Nishimasu H, Ran FA, Hsu PD, et al. Crystal structure of Cas9 in complex with guide RNA and target DNA. Cell. 2014;156:935‐949. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Kim E, Koo T, Park SW, et al. In vivo genome editing with a small Cas9 orthologue derived from campylobacter jejuni. Nat Commun. 2017;8:14500. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Ran FA, Hsu PD, Lin CY, et al. Double nicking by RNA‐guided CRISPR Cas9 for enhanced genome editing specificity. Cell. 2013;154:1380‐1389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Shen B, Zhang W, Zhang J, et al. Efficient genome modification by CRISPR‐Cas9 nickase with minimal off‐target effects. Nat Methods. 2014;11:399‐402. [DOI] [PubMed] [Google Scholar]
- 41. Qi LS, Larson MH, Gilbert LA, et al. Repurposing CRISPR as an RNA‐guided platform for sequence‐specific control of gene expression. Cell. 2013;152:1173‐1183. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Liao HK, Hatanaka F, Araoka T, et al. In vivo target gene activation via CRISPR/Cas9‐mediated trans‐epigenetic modulation. Cell. 2017;171:1495‐507.e15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Jensen TI, Mikkelsen NS, Gao Z, et al. Targeted regulation of transcription in primary cells using CRISPRa and CRISPRi. Genome Res. 2021;31:2120‐2130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Nuñez JK, Chen J, Pommier GC, et al. Genome‐wide programmable transcriptional memory by CRISPR‐based epigenome editing. Cell. 2021;184:2503‐19.e17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Walton RT, Christie KA, Whittaker MN, et al. Unconstrained genome targeting with near‐PAMless engineered CRISPR‐Cas9 variants. Science. 2020;368:290‐296. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Dong D, Ren K, Qiu X, et al. The crystal structure of Cpf1 in complex with CRISPR RNA. Nature. 2016;532:522‐526. [DOI] [PubMed] [Google Scholar]
- 47. Lei C, Li SY, Liu JK, et al. The CCTL (Cpf1‐assisted cutting and Taq DNA ligase‐assisted ligation) method for efficient editing of large DNA constructs in vitro. Nucleic Acids Res. 2017;45:e74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. Zetsche B, Gootenberg JS, Abudayyeh OO, et al. Cpf1 is a single RNA‐guided endonuclease of a class 2 CRISPR‐Cas system. Cell. 2015;163:759‐771. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Kim HK, Song M, Lee J, et al. In vivo high‐throughput profiling of CRISPR‐Cpf1 activity. Nat Methods. 2017;14:153‐159. [DOI] [PubMed] [Google Scholar]
- 50. Gao L, Cox DBT, Yan WX, et al. Engineered Cpf1 variants with altered PAM specificities. Nat Biotechnol. 2017;35:789‐792. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Kleinstiver BP, Sousa AA, Walton RT, et al. Engineered CRISPR‐Cas12a variants with increased activities and improved targeting ranges for gene, epigenetic and base editing. Nat Biotechnol. 2019;37:276‐282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Tu M, Lin L, Cheng Y, et al. A ‘new lease of life’: FnCpf1 possesses DNA cleavage activity for genome editing in human cells. Nucleic Acids Res. 2017;45:11295‐11304. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Kim D, Kim J, Hur JK, et al. Genome‐wide analysis reveals specificities of Cpf1 endonucleases in human cells. Nat Biotechnol. 2016;34:863‐868. [DOI] [PubMed] [Google Scholar]
- 54. Kleinstiver BP, Tsai SQ, Prew MS, et al. Genome‐wide specificities of CRISPR‐Cas Cpf1 nucleases in human cells. Nat Biotechnol. 2016;34:869‐874. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. O'Geen H, Ren C, Nicolet CM, et al. dCas9‐based epigenome editing suggests acquisition of histone methylation is not sufficient for target gene repression. Nucleic Acids Res. 2017;45:9901‐9916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Safari F, Zare K, Negahdaripour M, et al. CRISPR Cpf1 proteins: structure, function and implications for genome editing. Cell Biosci. 2019;9:36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57. Morita S, Horii T, Kimura M, et al. Synergistic upregulation of target genes by TET1 and VP64 in the dCas9‐SunTag platform. Int J Mol Sci. 2020;21:1574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Zhang X, Wang W, Shan L, et al. Gene activation in human cells using CRISPR/Cpf1‐p300 and CRISPR/Cpf1‐SunTag systems. Protein Cell. 2018;9:380‐383. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Burstein D, Harrington LB, Strutt SC, et al. New CRISPR‐Cas systems from uncultivated microbes. Nature. 2017;542:237‐241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Harrington LB, Burstein D, Chen JS, et al. Programmed DNA destruction by miniature CRISPR‐Cas14 enzymes. Science. 2018;362:839‐842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Yan WX, Hunnewell P, Alfonse LE, et al. Functionally diverse type V CRISPR‐Cas systems. Science. 2019;363:88‐91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Pausch P, Al‐Shayeb B, Bisom‐Rapp E, et al. CRISPR‐CasΦ from huge phages is a hypercompact genome editor. Science. 2020;369:333‐337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Liu L, Li X, Wang J, et al. Two distant catalytic sites are responsible for C2c2 RNase activities. Cell. 2017;168:121‐34.e12. [DOI] [PubMed] [Google Scholar]
- 64. Liu L, Li X, Ma J, et al. The molecular architecture for RNA‐guided RNA cleavage by Cas13a. Cell. 2017;170:714‐26.e10. [DOI] [PubMed] [Google Scholar]
- 65. Knott GJ, East‐Seletsky A, Cofsky JC, et al. Guide‐bound structures of an RNA‐targeting A‐cleaving CRISPR‐Cas13a enzyme. Nat Struct Mol Biol. 2017;24:825‐833. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. O'Connell MR. Molecular mechanisms of RNA targeting by Cas13‐containing type VI CRISPR‐Cas systems. J Mol Biol. 2019;431:66‐87. [DOI] [PubMed] [Google Scholar]
- 67. Yan F, Wang W, Zhang J. CRISPR‐Cas12 and Cas13: the lesser known siblings of CRISPR‐Cas9. Cell Biol Toxicol. 2019;35:489‐492. [DOI] [PubMed] [Google Scholar]
- 68. Slaymaker IM, Mesa P, Kellner MJ, et al. High‐resolution structure of Cas13b and biochemical characterization of RNA targeting and cleavage. Cell Rep. 2019;26:3741‐51.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Zhang C, Konermann S, Brideau NJ, et al. Structural basis for the RNA‐guided ribonuclease activity of CRISPR‐Cas13d. Cell. 2018;175:212‐23.e17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Abudayyeh OO, Gootenberg JS, Konermann S, et al. C2c2 is a single‐component programmable RNA‐guided RNA‐targeting CRISPR effector. Science. 2016;353:aaf5573. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71. Pickar‐Oliver A, Gersbach CA. The next generation of CRISPR‐Cas technologies and applications. Nat Rev Mol Cell Biol. 2019;20:490‐507. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Collias D, Beisel CL. CRISPR technologies and the search for the PAM‐free nuclease. Nat Commun. 2021;12:555. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. East‐Seletsky A, O'Connell MR, Knight SC, et al. Two distinct RNase activities of CRISPR‐C2c2 enable guide‐RNA processing and RNA detection. Nature. 2016;538:270‐273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74. Konermann S, Lotfy P, Brideau NJ, et al. Transcriptome engineering with RNA‐targeting type VI‐D CRISPR effectors. Cell. 2018;173:665‐76.e14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Du M, Jillette N, Zhu JJ, et al. CRISPR artificial splicing factors. Nat Commun. 2020;11:2973. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Kordyś M, Sen R, Warkocki Z. Applications of the versatile CRISPR‐Cas13 RNA targeting system. Wiley Interdiscip Rev RNA. 2022;13:e1694. doi: 10.1002/wrna.1694.e1694 [DOI] [PubMed] [Google Scholar]
- 77. Xu D, Cai Y, Tang L, et al. A CRISPR/Cas13‐based approach demonstrates biological relevance of vlinc class of long non‐coding RNAs in anticancer drug response. Sci Rep. 2020;10:1794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Li J, Chen Z, Chen F, et al. Targeted mRNA demethylation using an engineered dCas13b‐ALKBH5 fusion protein. Nucleic Acids Res. 2020;48:5684‐5694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Khalil AS, Collins JJ. Synthetic biology: applications come of age. Nat Rev Genet. 2010;11:367‐379. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80. Hutchison CA 3rd, Chuang RY, Noskov VN, et al. Design and synthesis of a minimal bacterial genome. Science. 2016;351:aad6253. [DOI] [PubMed] [Google Scholar]
- 81. Pelletier JF, Sun L, Wise KS, et al. Genetic requirements for cell division in a genomically minimal cell. Cell. 2021;184:2430‐40.e16. [DOI] [PubMed] [Google Scholar]
- 82. Quinn JY, Cox RS 3rd, Adler A, et al. SBOL visual: a graphical language for genetic designs. PLoS Biol. 2015;13:e1002310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83. Mishra D, Rivera PM, Lin A, et al. A load driver device for engineering modularity in biological networks. Nat Biotechnol. 2014;32:1268‐1275. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Nielsen AA, Der BS, Shin J, et al. Genetic circuit design automation. Science. 2016;352:aac7341. [DOI] [PubMed] [Google Scholar]
- 85. Gardner TS, Cantor CR, Collins JJ. Construction of a genetic toggle switch in Escherichia coli. Nature. 2000;403:339‐342. [DOI] [PubMed] [Google Scholar]
- 86. Tan X, Letendre JH, Collins JJ, et al. Synthetic biology in the clinic: engineering vaccines, diagnostics, and therapeutics. Cell. 2021;184:881‐898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Williams JZ, Allen GM, Shah D, et al. Precise T cell recognition programs designed by transcriptionally linking multiple receptors. Science. 2020;370:1099‐1104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88. Wong AS, Choi GC, Cui CH, et al. Multiplexed barcoded CRISPR‐Cas9 screening enabled by CombiGEM. Proc Natl Acad Sci USA. 2016;113:2544‐2549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89. Dong C, Fontana J, Patel A, et al. Synthetic CRISPR‐Cas gene activators for transcriptional reprogramming in bacteria. Nat Commun. 2018;9:2489. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Santos‐Moreno J, Schaerli Y. CRISPR‐based gene expression control for synthetic gene circuits. Biochem Soc Trans. 2020;48:1979‐1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Gander MW, Vrana JD, Voje WE, et al. Digital logic circuits in yeast with CRISPR‐dCas9 NOR gates. Nat Commun. 2017;8:15459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Hofmann A, Falk J, Prangemeier T, et al. A tightly regulated and adjustable CRISPR‐dCas9 based AND gate in yeast. Nucleic Acids Res. 2019;47:509‐520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Peng L, Zhou J, Yin L, et al. Integration of logic gates to CRISPR/Cas12a system for rapid and sensitive detection of pathogenic bacterial genes. Anal Chim Acta. 2020;1125:162‐168. [DOI] [PubMed] [Google Scholar]
- 94. Kempton HR, Goudy LE, Love KS, et al. Multiple input sensing and signal integration using a Split Cas12a system. Mol Cell. 2020;78:184‐91.e3. [DOI] [PubMed] [Google Scholar]
- 95. Stahler A, Stintzing S, von Einem JC, et al. Single‐nucleotide variants, tumour mutational burden and microsatellite instability in patients with metastatic colorectal cancer: next‐generation sequencing results of the FIRE‐3 trial. Eur J Cancer. 2020;137:250‐259. [DOI] [PubMed] [Google Scholar]
- 96. Chang W, Liu W, Shen H, et al. Molecular AND logic gate for multiple single‐nucleotide mutations detection based on CRISPR/Cas9n system‐trigged signal amplification. Anal Chim Acta. 2020;1112:46‐53. [DOI] [PubMed] [Google Scholar]
- 97. Leão R, Apolónio JD, Lee D, et al. Mechanisms of human telomerase reverse transcriptase (hTERT) regulation: clinical impacts in cancer. J Biomed Sci. 2018;25:22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Li J, Huang C, Xiong T, et al. A CRISPR interference of CBP and p300 selectively induced synthetic lethality in bladder cancer cells In vitro. Int J Biol Sci. 2019;15:1276‐1286. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99. Liu Y, Zeng Y, Liu L, et al. Synthesizing AND gate genetic circuits based on CRISPR‐Cas9 for identification of bladder cancer cells. Nat Commun. 2014;5:5393. [DOI] [PubMed] [Google Scholar]
- 100. Liu Y, Huang W, Cai Z. Synthesizing AND gate minigene circuits based on CRISPReader for identification of bladder cancer cells. Nat Commun. 2020;11:5486. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Ye H, Fussenegger M. Optogenetic medicine: synthetic therapeutic solutions precision‐guided by light. Cold Spring Harb Perspect Med. 2019;9(9):a034371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Mansouri M, Strittmatter T, Fussenegger M. Light‐controlled mammalian cells and their therapeutic applications in synthetic biology. Adv Sci. 2019;6:1800952. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103. Polstein LR, Gersbach CA. A light‐inducible CRISPR‐Cas9 system for control of endogenous gene activation. Nat Chem Biol. 2015;11:198‐200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Amrani N, Gao XD, Liu P, et al. NmeCas9 is an intrinsically high‐fidelity genome‐editing platform. Genome Biol. 2018;19:214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105. Hoffmann MD, Mathony J, Upmeier Zu Belzen J, et al. Optogenetic control of Neisseria meningitidis Cas9 genome editing using an engineered, light‐switchable anti‐CRISPR protein. Nucleic Acids Res. 2021;49:e29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106. Qi F, Tan B, Ma F, et al. A synthetic light‐switchable system based on CRISPR Cas13a regulates the expression of LncRNA MALAT1 and affects the malignant phenotype of bladder cancer cells. Int J Biol Sci. 2019;15:1630‐1636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Ash C, Dubec M, Donne K, et al. Effect of wavelength and beam width on penetration in light‐tissue interaction using computational methods. Lasers Med Sci. 2017;32:1909‐1918. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Nakashima Y, Ohta S, Wolf AM. Blue light‐induced oxidative stress in live skin. Free Radic Biol Med. 2017;108:300‐310. [DOI] [PubMed] [Google Scholar]
- 109. Chen X, Chen Y, Xin H, et al. Near‐infrared optogenetic engineering of photothermal nanoCRISPR for programmable genome editing. Proc Natl Acad Sci USA. 2020;117:2395‐2405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110. Pan Y, Yang J, Luan X, et al. Near‐infrared upconversion‐activated CRISPR‐Cas9 system: a remote‐controlled gene editing platform. Sci Adv. 2019;5:eaav7199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111. AlAjmi MF, Rehman MT, Hussain A, et al. Pharmacoinformatics approach for the identification of polo‐like kinase‐1 inhibitors from natural sources as anti‐cancer agents. Int J Biol Macromol. 2018;116:173‐181. [DOI] [PubMed] [Google Scholar]
- 112. Shakil S, Baig MH, Tabrez S, et al. Molecular and enzoinformatics perspectives of targeting polo‐like kinase 1 in cancer therapy. Semin Cancer Biol. 2019;56:47‐55. [DOI] [PubMed] [Google Scholar]
- 113. Yu Y, Wu X, Guan N, et al. Engineering a far‐red light‐activated split‐Cas9 system for remote‐controlled genome editing of internal organs and tumors. Sci Adv. 2020;6:eabb1777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. Kebschull JM, Zador AM. Cellular barcoding: lineage tracing, screening and beyond. Nat Methods. 2018;15:871‐879. [DOI] [PubMed] [Google Scholar]
- 115. Rodriguez‐Fraticelli AE, Wolock SL, Weinreb CS, et al. Clonal analysis of lineage fate in native haematopoiesis. Nature. 2018;553:212‐216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Merino D, Weber TS, Serrano A, et al. Barcoding reveals complex clonal behavior in patient‐derived xenografts of metastatic triple negative breast cancer. Nat Commun. 2019;10:766. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117. Spanjaard B, Hu B, Mitic N, et al. Simultaneous lineage tracing and cell‐type identification using CRISPR‐Cas9‐induced genetic scars. Nat Biotechnol. 2018;36:469‐473. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118. Kalhor R, Mali P, Church GM. Rapidly evolving homing CRISPR barcodes. Nat Methods. 2017;14:195‐200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Kalhor R, Kalhor K, Mejia L, et al. Developmental barcoding of whole mouse via homing CRISPR. Science. 2018;361(6405):eaat9804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120. McKenna A, Findlay GM, Gagnon JA, et al. Whole‐organism lineage tracing by combinatorial and cumulative genome editing. Science. 2016;353:aaf7907. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121. Raj B, Wagner DE, McKenna A, et al. Simultaneous single‐cell profiling of lineages and cell types in the vertebrate brain. Nat Biotechnol. 2018;36:442‐450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122. Raj B, Gagnon JA, Schier AF. Large‐scale reconstruction of cell lineages using single‐cell readout of transcriptomes and CRISPR‐Cas9 barcodes by scGESTALT. Nat Protoc. 2018;13:2685‐2713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123. Bowling S, Sritharan D, Osorio FG, et al. An engineered CRISPR‐Cas9 mouse line for simultaneous readout of lineage histories and gene expression profiles in single cells. Cell. 2020;181:1410‐22.e27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124. Zhang W, Bado IL, Hu J, et al. The bone microenvironment invigorates metastatic seeds for further dissemination. Cell. 2021;184:2471‐86.e20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Rogers ZN, McFarland CD, Winters IP, et al. Mapping the in vivo fitness landscape of lung adenocarcinoma tumor suppression in mice. Nat Genet. 2018;50:483‐486. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126. Shalem O, Sanjana NE, Zhang F. High‐throughput functional genomics using CRISPR‐Cas9. Nat Rev Genet. 2015;16:299‐311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127. Wroblewska A, Dhainaut M, Ben‐Zvi B, et al. Protein barcodes enable High‐dimensional single‐cell CRISPR screens. Cell. 2018;175:1141‐55.e16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Ma Y, Zhu Q, Wang X, et al. Synthetic lethal screening identifies DHODH as a target for MEN1‐mutated tumor cells. Cell Res. 2022;32:596‐599. doi: 10.1038/s41422-022-00613-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129. Qian Y, Liang X, Kong P, et al. Elevated DHODH expression promotes cell proliferation via stabilizing β‐catenin in esophageal squamous cell carcinoma. Cell Death Dis. 2020;11:862. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130. Kaur H, Sarma P, Bhattacharyya A, et al. Efficacy and safety of dihydroorotate dehydrogenase (DHODH) inhibitors "leflunomide" and "teriflunomide" in Covid‐19: a narrative review. Eur J Pharmacol. 2021;906:174233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Steinhart Z, Pavlovic Z, Chandrashekhar M, et al. Genome‐wide CRISPR screens reveal a Wnt‐FZD5 signaling circuit as a druggable vulnerability of RNF43‐mutant pancreatic tumors. Nat Med. 2017;23:60‐68. [DOI] [PubMed] [Google Scholar]
- 132. Lawson KA, Sousa CM, Zhang X, et al. Functional genomic landscape of cancer‐intrinsic evasion of killing by T cells. Nature. 2020;586:120‐126. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133. Dixit A, Parnas O, Li B, et al. Perturb‐Seq: dissecting molecular circuits with scalable single‐cell RNA profiling of pooled genetic screens. Cell. 2016;167:1853‐66.e17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134. Jaitin DA, Weiner A, Yofe I, et al. Dissecting immune circuits by linking CRISPR‐pooled screens with single‐cell RNA‐Seq. Cell. 2016;167:1883‐96.e15. [DOI] [PubMed] [Google Scholar]
- 135. Datlinger P, Rendeiro AF, Schmidl C, et al. Pooled CRISPR screening with single‐cell transcriptome readout. Nat Methods. 2017;14:297‐301. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136. Wheeler EC, Vu AQ, Einstein JM, et al. Pooled CRISPR screens with imaging on microraft arrays reveals stress granule‐regulatory factors. Nat Methods. 2020;17:636‐642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137. Ringel T, Frey N, Ringnalda F, et al. Genome‐scale CRISPR screening in human intestinal organoids identifies drivers of TGF‐β resistance. Cell Stem Cell. 2020;26:431‐40.e8. [DOI] [PubMed] [Google Scholar]
- 138. Michels BE, Mosa MH, Streibl BI, et al. Pooled In vitro and In vivo CRISPR‐Cas9 screening identifies tumor suppressors in human colon organoids. Cell Stem Cell. 2020;26:782‐92.e7. [DOI] [PubMed] [Google Scholar]
- 139. Kim S, Kim D, Cho SW, et al. Highly efficient RNA‐guided genome editing in human cells via delivery of purified Cas9 ribonucleoproteins. Genome Res. 2014;24:1012‐1019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140. Woo JW, Kim J, Kwon SI, et al. DNA‐free genome editing in plants with preassembled CRISPR‐Cas9 ribonucleoproteins. Nat Biotechnol. 2015;33:1162‐1164. [DOI] [PubMed] [Google Scholar]
- 141. Niu Y, Shen B, Cui Y, et al. Generation of gene‐modified cynomolgus monkey via Cas9/RNA‐mediated gene targeting in one‐cell embryos. Cell. 2014;156:836‐843. [DOI] [PubMed] [Google Scholar]
- 142. Ran FA, Hsu PD, Wright J, et al. Genome engineering using the CRISPR‐Cas9 system. Nat Protoc. 2013;8:2281‐2308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143. Xu X, Liu C, Wang Y, et al. Nanotechnology‐based delivery of CRISPR/Cas9 for cancer treatment. Adv Drug Deliv Rev. 2021;176:113891. [DOI] [PubMed] [Google Scholar]
- 144. Xu CL, Ruan MZC, Mahajan VB, et al. Viral delivery systems for CRISPR. Viruses. 2019;11(1):28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145. Xu CF, Chen GJ, Luo YL, et al. Rational designs of in vivo CRISPR‐Cas delivery systems. Adv Drug Deliv Rev. 2021;168:3‐29. [DOI] [PubMed] [Google Scholar]
- 146. Yeh TK, Jean SS, Lee YL, et al. Bacteriophages and phage‐delivered CRISPR‐Cas system as antibacterial therapy. Int J Antimicrob Agents. 2022;59:106475. [DOI] [PubMed] [Google Scholar]
- 147. Mingozzi F, High KA. Therapeutic in vivo gene transfer for genetic disease using AAV: progress and challenges. Nat Rev Genet. 2011;12:341‐355. [DOI] [PubMed] [Google Scholar]
- 148. Grieger JC, Samulski RJ. Packaging capacity of adeno‐associated virus serotypes: impact of larger genomes on infectivity and postentry steps. J Virol. 2005;79:9933‐9944. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149. Wang D, Zhang F, Gao G. CRISPR‐based therapeutic genome editing: strategies and In vivo delivery by AAV vectors. Cell. 2020;181:136‐150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150. Murovec J, Pirc Ž, Yang B. New variants of CRISPR RNA‐guided genome editing enzymes. Plant Biotechnol J. 2017;15:917‐926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151. Yang Y, Wang L, Bell P, et al. A dual AAV system enables the Cas9‐mediated correction of a metabolic liver disease in newborn mice. Nat Biotechnol. 2016;34:334‐338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152. Richards DY, Winn SR, Dudley S, et al. AAV‐mediated CRISPR/Cas9 gene editing in murine phenylketonuria. Mol Ther Methods Clin Dev. 2020;17:234‐245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153. Aubert M, Strongin DE, Roychoudhury P, et al. Gene editing and elimination of latent herpes simplex virus in vivo. Nat Commun. 2020;11:4148. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154. Lee CS, Bishop ES, Zhang R, et al. Adenovirus‐mediated gene delivery: potential applications for gene and Cell‐based therapies in the new era of personalized medicine. Genes Dis. 2017;4:43‐63. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155. Ricobaraza A, Gonzalez‐Aparicio M, Mora‐Jimenez L, et al. High‐capacity adenoviral vectors: expanding the scope of gene therapy. Int J Mol Sci. 2020;21(10):3643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156. Kato Y, Tabata H, Sato K, et al. Adenovirus vectors expressing eight multiplex guide RNAs of CRISPR/Cas9 efficiently disrupted diverse hepatitis B virus gene derived from heterogeneous patient. Int J Mol Sci. 2021;22(19):10570. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157. Collignon C, Bol V, Chalon A, et al. Innate immune responses to chimpanzee adenovirus vector 155 vaccination in mice and monkeys. Front Immunol. 2020;11:579872. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158. Collins LT, Curiel DT. Synthetic biology approaches for engineering next‐generation adenoviral gene therapies. ACS Nano. 2021;15:13970‐13979. [DOI] [PubMed] [Google Scholar]
- 159. Magrin E, Miccio A, Cavazzana M. Lentiviral and genome‐editing strategies for the treatment of β‐hemoglobinopathies. Blood. 2019;134:1203‐1213. [DOI] [PubMed] [Google Scholar]
- 160. Page A, Fusil F, Cosset FL. Toward tightly tuned gene expression following lentiviral vector transduction. Viruses. 2020;12(12):1427. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161. Pauwels K, Gijsbers R, Toelen J, et al. State‐of‐the‐art lentiviral vectors for research use: risk assessment and biosafety recommendations. Curr Gene Ther. 2009;9:459‐474. [DOI] [PubMed] [Google Scholar]
- 162. Kotterman MA, Chalberg TW, Schaffer DV. Viral vectors for gene therapy: translational and clinical outlook. Annu Rev Biomed Eng. 2015;17:63‐89. [DOI] [PubMed] [Google Scholar]
- 163. Holmgaard A, Askou AL, Benckendorff JNE, et al. In vivo knockout of the Vegfa gene by lentiviral delivery of CRISPR/Cas9 in mouse retinal pigment epithelium cells. Mol Ther Nucleic Acids. 2017;9:89‐99. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164. Aubrey BJ, Kelly GL, Kueh AJ, et al. An inducible lentiviral guide RNA platform enables the identification of tumor‐essential genes and tumor‐promoting mutations in vivo. Cell Rep. 2015;10:1422‐1432. [DOI] [PubMed] [Google Scholar]
- 165. Biasco L, Rothe M, Büning H, et al. Analyzing the genotoxicity of retroviral vectors in hematopoietic cell gene therapy. Mol Ther Methods Clin Dev. 2018;8:21‐30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166. Poletti V, Mavilio F. Interactions between retroviruses and the host cell genome. Mol Ther Methods Clin Dev. 2018;8:31‐41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167. Bikard D, Euler CW, Jiang W, et al. Exploiting CRISPR‐Cas nucleases to produce sequence‐specific antimicrobials. Nat Biotechnol. 2014;32:1146‐1150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168. Selle K, Fletcher JR, Tuson H, et al. In vivo targeting of Clostridioides difficile using phage‐delivered CRISPR‐Cas3 antimicrobials. MBio. 2020;11(2):e00019‐e00020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169. Liu H, Li H, Liang Y, et al. Phage‐delivered sensitisation with subsequent antibiotic treatment reveals sustained effect against antimicrobial resistant bacteria. Theranostics. 2020;10:6310‐6321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170. Li J, Røise JJ, He M, et al. Non‐viral strategies for delivering genome editing enzymes. Adv Drug Deliv Rev. 2021;168:99‐117. [DOI] [PubMed] [Google Scholar]
- 171. Salimi M, Shirazi A, Sineh Sepehr K, et al. The effect of CRISPR constructs microinjection on the expression of developmental genes in Rag1 knocked‐out mice embryo. Vet Med Sci. 2021;7:1436‐1442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172. Liang P, Xu Y, Zhang X, et al. CRISPR/Cas9‐mediated gene editing in human tripronuclear zygotes. Protein Cell. 2015;6:363‐372. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173. Ai D, Wang B, Fan Z, et al. Embryo microinjection and knockout mutant identification of CRISPR/Cas9 genome‐edited Helicoverpa Armigera (Hübner). J Vis Exp. 2021;173:e62068. doi: 10.3791/62068 [DOI] [PubMed] [Google Scholar]
- 174. Han X, Liu Z, Jo MC, et al. CRISPR‐Cas9 delivery to hard‐to‐transfect cells via membrane deformation. Sci Adv. 2015;1:e1500454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175. Nelson CE, Hakim CH, Ousterout DG, et al. In vivo genome editing improves muscle function in a mouse model of Duchenne muscular dystrophy. Science. 2016;351:403‐407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176. Chen S, Lee B, Lee AY, et al. Highly efficient mouse genome editing by CRISPR ribonucleoprotein electroporation of zygotes. J Biol Chem. 2016;291:14457‐14467. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177. Cui H, Jiang M, Zhou W, et al. Carrier‐free cellular transport of CRISPR/Cas9 ribonucleoprotein for genome editing by cold atmospheric plasma. Biology. 2021;10(10):1038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178. Xu Z, Hu J, Cao H, et al. Loss of Pten synergizes with c‐met to promote hepatocellular carcinoma development via mTORC2 pathway. Exp Mol Med. 2018;50:e417. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179. Wang HX, Li M, Lee CM, et al. CRISPR/Cas9‐based genome editing for disease modeling and therapy: challenges and opportunities for nonviral delivery. Chem Rev. 2017;117:9874‐9906. [DOI] [PubMed] [Google Scholar]
- 180. Yu Q, Tan RZ, Gan Q, et al. A novel rat model of nonalcoholic fatty liver disease constructed through CRISPR/Cas‐based hydrodynamic injection. Mol Biotechnol. 2017;59:365‐373. [DOI] [PubMed] [Google Scholar]
- 181. Sessions JW, Skousen CS, Price KD, et al. CRISPR‐Cas9 directed knock‐out of a constitutively expressed gene using lance array nanoinjection. Springerplus. 2016;5:1521. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182. Esteban‐Fernández de Ávila B, Angell C, Soto F, et al. Acoustically propelled Nanomotors for intracellular siRNA delivery. ACS Nano. 2016;10:4997‐5005. [DOI] [PubMed] [Google Scholar]
- 183. Hansen‐Bruhn M, de Ávila BE, Beltrán‐Gastélum M, et al. Active intracellular delivery of a Cas9/sgRNA complex using ultrasound‐propelled Nanomotors. Angew Chem Int Ed Engl. 2018;57:2657‐2661. [DOI] [PubMed] [Google Scholar]
- 184. Liu S, Cheng Q, Wei T, et al. Membrane‐destabilizing ionizable phospholipids for organ‐selective mRNA delivery and CRISPR‐Cas gene editing. Nat Mater. 2021;20:701‐710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185. Khan MM, Filipczak N, Torchilin VP. Cell penetrating peptides: a versatile vector for co‐delivery of drug and genes in cancer. J Control Release. 2021;330:1220‐1228. [DOI] [PubMed] [Google Scholar]
- 186. Tang H, Zhao X, Jiang X. Synthetic multi‐layer nanoparticles for CRISPR‐Cas9 genome editing. Adv Drug Deliv Rev. 2021;168:55‐78. [DOI] [PubMed] [Google Scholar]
- 187. Liu C, Zhang L, Liu H, et al. Delivery strategies of the CRISPR‐Cas9 gene‐editing system for therapeutic applications. J Control Release. 2017;266:17‐26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188. Chen F, Alphonse M, Liu Q. Strategies for nonviral nanoparticle‐based delivery of CRISPR/Cas9 therapeutics. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2020;12:e1609. [DOI] [PubMed] [Google Scholar]
- 189. Zuris JA, Thompson DB, Shu Y, et al. Cationic lipid‐mediated delivery of proteins enables efficient protein‐based genome editing in vitro and in vivo. Nat Biotechnol. 2015;33:73‐80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190. Cho EY, Ryu JY, Lee HAR, et al. Lecithin nano‐liposomal particle as a CRISPR/Cas9 complex delivery system for treating type 2 diabetes. J Nanobiotechnol. 2019;17:19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191. Zhang D, Wang G, Yu X, et al. Enhancing CRISPR/Cas gene editing through modulating cellular mechanical properties for cancer therapy. Nat Nanotechnol. 2022;17:777‐787. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192. Yin H, Song CQ, Dorkin JR, et al. Therapeutic genome editing by combined viral and non‐viral delivery of CRISPR system components in vivo. Nat Biotechnol. 2016;34:328‐333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Ryu N, Kim MA, Park D, et al. Effective PEI‐mediated delivery of CRISPR‐Cas9 complex for targeted gene therapy. Nanomedicine. 2018;14:2095‐2102. [DOI] [PubMed] [Google Scholar]
- 194. Luo YL, Xu CF, Li HJ, et al. Macrophage‐specific in vivo gene editing using cationic lipid‐assisted polymeric nanoparticles. ACS Nano. 2018;12:994‐1005. [DOI] [PubMed] [Google Scholar]
- 195. Trofimenko E, Grasso G, Heulot M, et al. Genetic, cellular, and structural characterization of the membrane potential‐dependent cell‐penetrating peptide translocation pore. Elife. 2021;10:e69832. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196. Suresh B, Ramakrishna S, Kim H. Cell‐penetrating peptide‐mediated delivery of Cas9 protein and guide RNA for genome editing. Methods Mol Biol. 2017;1507:81‐94. [DOI] [PubMed] [Google Scholar]
- 197. Ramakrishna S, Kwaku Dad AB, Beloor J, et al. Gene disruption by cell‐penetrating peptide‐mediated delivery of Cas9 protein and guide RNA. Genome Res. 2014;24:1020‐1027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198. Lostalé‐Seijo I, Louzao I, Juanes M, et al. Peptide/Cas9 nanostructures for ribonucleoprotein cell membrane transport and gene edition. Chem Sci. 2017;8:7923‐7931. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199. Kardani K, Milani A, Shabani HS, et al. Cell penetrating peptides: the potent multi‐cargo intracellular carriers. Expert Opin Drug Deliv. 2019;16:1227‐1258. [DOI] [PubMed] [Google Scholar]
- 200. Fan J, Cheng Y, Sun M. Functionalized gold nanoparticles: synthesis, properties and biomedical applications. Chem Rec. 2020;20:1474‐1504. [DOI] [PubMed] [Google Scholar]
- 201. Shrestha B, Wang L, Zhang H, et al. Gold nanoparticles mediated drug‐gene combinational therapy for breast cancer treatment. Int J Nanomedicine. 2020;15:8109‐8119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202. Mout R, Ray M, Yesilbag Tonga G, et al. Direct cytosolic delivery of CRISPR/Cas9‐ribonucleoprotein for efficient gene editing. ACS Nano. 2017;11:2452‐2458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203. Wang P, Zhang L, Zheng W, et al. Thermo‐triggered release of CRISPR‐Cas9 system by lipid‐encapsulated gold nanoparticles for tumor therapy. Angew Chem Int Ed Engl. 2018;57:1491‐1496. [DOI] [PubMed] [Google Scholar]
- 204. Tao Y, Yi K, Hu H, et al. Coassembly of nucleus‐targeting gold nanoclusters with CRISPR/Cas9 for simultaneous bioimaging and therapeutic genome editing. J Mater Chem B. 2021;9:94‐100. [DOI] [PubMed] [Google Scholar]
- 205. Yue H, Zhou X, Cheng M, et al. Graphene oxide‐mediated Cas9/sgRNA delivery for efficient genome editing. Nanoscale. 2018;10:1063‐1071. [DOI] [PubMed] [Google Scholar]
- 206. Zhou W, Cui H, Ying L, et al. Enhanced cytosolic delivery and release of CRISPR/Cas9 by Black phosphorus Nanosheets for genome editing. Angew Chem Int Ed Engl. 2018;57:10268‐10272. [DOI] [PubMed] [Google Scholar]
- 207. Li M, Wang C, Di Z, et al. Engineering multifunctional DNA hybrid Nanospheres through coordination‐driven self‐assembly. Angew Chem Int Ed Engl. 2019;58:1350‐1354. [DOI] [PubMed] [Google Scholar]
- 208. Mathew B, Ramanathan R, Delvaux NA, et al. Heat‐shrinking DNA nanoparticles for in vivo gene delivery. Gene Ther. 2020;27:196‐208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209. Han S, Lee Y, Lee M. Biomimetic cell membrane‐coated DNA nanoparticles for gene delivery to glioblastoma. J Control Release. 2021;338:22‐32. [DOI] [PubMed] [Google Scholar]
- 210. Sun W, Ji W, Hall JM, et al. Self‐assembled DNA nanoclews for the efficient delivery of CRISPR‐Cas9 for genome editing. Angew Chem Int Ed Engl. 2015;54:12029‐12033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211. O'Brien K, Breyne K, Ughetto S, et al. RNA delivery by extracellular vesicles in mammalian cells and its applications. Nat Rev Mol Cell Biol. 2020;21:585‐606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212. Zubarev I, Vladimirtsev D, Vorontsova M, et al. Viral membrane fusion proteins and RNA sorting mechanisms for the molecular delivery by exosomes. Cell. 2021;10(11):3043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213. Yip BH. Recent advances in CRISPR/Cas9 delivery strategies. Biomolecules. 2020;10(6):839. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214. Zhuang J, Tan J, Wu C, et al. Extracellular vesicles engineered with valency‐controlled DNA nanostructures deliver CRISPR/Cas9 system for gene therapy. Nucleic Acids Res. 2020;48:8870‐8882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215. Yao X, Lyu P, Yoo K, et al. Engineered extracellular vesicles as versatile ribonucleoprotein delivery vehicles for efficient and safe CRISPR genome editing. J Extracell Vesicles. 2021;10:e12076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216. Campbell LA, Coke LM, Richie CT, et al. Gesicle‐mediated delivery of CRISPR/Cas9 ribonucleoprotein complex for inactivating the HIV provirus. Mol Ther. 2019;27:151‐163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217. Gillmore JD, Gane E, Taubel J, et al. CRISPR‐Cas9 In vivo gene editing for transthyretin amyloidosis. N Engl J Med. 2021;385:493‐502. [DOI] [PubMed] [Google Scholar]
- 218. Lu Y, Xue J, Deng T, et al. Safety and feasibility of CRISPR‐edited T cells in patients with refractory non‐small‐cell lung cancer. Nat Med. 2020;26:732‐740. [DOI] [PubMed] [Google Scholar]
- 219. Yin H, Xue W, Anderson DG. CRISPR‐Cas: a tool for cancer research and therapeutics. Nat Rev Clin Oncol. 2019;16:281‐295. [DOI] [PubMed] [Google Scholar]
- 220. Heckl D, Kowalczyk MS, Yudovich D, et al. Generation of mouse models of myeloid malignancy with combinatorial genetic lesions using CRISPR‐Cas9 genome editing. Nat Biotechnol. 2014;32:941‐946. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221. Chen CY, Chen J, He L, et al. PTEN: tumor suppressor and metabolic regulator. Front Endocrinol. 2018;9:338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222. Meng X, Wang ZF, Lou QY, et al. Long non‐coding RNAs in head and neck squamous cell carcinoma: diagnostic biomarkers, targeted therapies, and prognostic roles. Eur J Pharmacol. 2021;902:174114. [DOI] [PubMed] [Google Scholar]
- 223. Xue W, Chen S, Yin H, et al. CRISPR‐mediated direct mutation of cancer genes in the mouse liver. Nature. 2014;514:380‐384. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224. Maresch R, Mueller S, Veltkamp C, et al. Multiplexed pancreatic genome engineering and cancer induction by transfection‐based CRISPR/Cas9 delivery in mice. Nat Commun. 2016;7:10770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225. Zhang Y, Zhang Z. The history and advances in cancer immunotherapy: understanding the characteristics of tumor‐infiltrating immune cells and their therapeutic implications. Cell Mol Immunol. 2020;17:807‐821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226. Manguso RT, Pope HW, Zimmer MD, et al. In vivo CRISPR screening identifies Ptpn2 as a cancer immunotherapy target. Nature. 2017;547:413‐418. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227. Dong MB, Wang G, Chow RD, et al. Systematic immunotherapy target discovery using genome‐scale In vivo CRISPR screens in CD8 T cells. Cell. 2019;178:1189‐204.e23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228. Wang X, Tokheim C, Gu SS, et al. In vivo CRISPR screens identify the E3 ligase Cop1 as a modulator of macrophage infiltration and cancer immunotherapy target. Cell. 2021;184:5357‐74.e22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229. Stadtmauer EA, Fraietta JA, Davis MM, et al. CRISPR‐engineered T cells in patients with refractory cancer. Science. 2020;367(6481):eaba7365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230. Hu Y, Zhou Y, Zhang M, et al. CRISPR/Cas9‐engineered universal CD19/CD22 dual‐targeted CAR‐T cell therapy for relapsed/refractory B‐cell acute lymphoblastic leukemia. Clin Cancer Res. 2021;27:2764‐2772. [DOI] [PubMed] [Google Scholar]
- 231. Hu W, Zi Z, Jin Y, et al. CRISPR/Cas9‐mediated PD‐1 disruption enhances human mesothelin‐targeted CAR T cell effector functions. Cancer Immunol Immunother. 2019;68:365‐377. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232. Chen M, Mao A, Xu M, et al. CRISPR‐Cas9 for cancer therapy: opportunities and challenges. Cancer Lett. 2019;447:48‐55. [DOI] [PubMed] [Google Scholar]
- 233. Zhang H, Qin C, An C, et al. Application of the CRISPR/Cas9‐based gene editing technique in basic research, diagnosis, and therapy of cancer. Mol Cancer. 2021;20:126. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234. Ihry RJ, Worringer KA, Salick MR, et al. p53 inhibits CRISPR‐Cas9 engineering in human pluripotent stem cells. Nat Med. 2018;24:939‐946. [DOI] [PubMed] [Google Scholar]
- 235. Huang X, Zhou Q, Wang M, et al. A light‐inducible Split‐dCas9 system for inhibiting the progression of bladder cancer cells by activating p53 and E‐cadherin. Front Mol Biosci. 2020;7:627848. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
The data that support the findings of this study are available from the corresponding author upon reasonable request.