Next Article in Journal
Coexistence of Genetic Diseases Is a New Clinical Challenge: Three Unrelated Cases of Dual Diagnosis
Previous Article in Journal
Independent Associated SNPs at SORCS3 and Its Protein Interactors for Multiple Brain-Related Disorders and Traits
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in and Prospects of Genome Editing Tools for Human Disease Model Development and Therapeutic Applications

1
Department of Physiology, Korea University College of Medicine, Seoul 02841, Republic of Korea
2
Department of Biomedical Sciences, Korea University College of Medicine, Seoul 02841, Republic of Korea
3
Graduate School of Cancer Science and Policy, National Cancer Center, Goyang 10408, Republic of Korea
4
Korea Mouse Phenotyping Center, Seoul National University, Seoul 08826, Republic of Korea
5
Laboratory of Developmental Biology and Genomics, BK21 PLUS Program for Creative Veterinary Science Research, Research Institute for Veterinary Science, College of Veterinary Medicine, Seoul National University, Seoul 08826, Republic of Korea
6
Interdisciplinary Program for Bioinformatics, Program for Cancer Biology, BIO-MAX/N-Bio Institute, Seoul National University, Seoul 08826, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2023, 14(2), 483; https://doi.org/10.3390/genes14020483
Submission received: 29 December 2022 / Revised: 9 February 2023 / Accepted: 10 February 2023 / Published: 14 February 2023
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

:
Programmable nucleases, such as zinc finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs), and clustered regularly interspaced short palindromic repeats (CRISPR)/Cas, are widely accepted because of their diversity and enormous potential for targeted genomic modifications in eukaryotes and other animals. Moreover, rapid advances in genome editing tools have accelerated the ability to produce various genetically modified animal models for studying human diseases. Given the advances in gene editing tools, these animal models are gradually evolving toward mimicking human diseases through the introduction of human pathogenic mutations in their genome rather than the conventional gene knockout. In the present review, we summarize the current progress in and discuss the prospects for developing mouse models of human diseases and their therapeutic applications based on advances in the study of programmable nucleases.

1. Introduction

Animal models are invaluable for unraveling pathogenic mechanisms and developing new therapeutics [1,2]. Given the long-standing efforts to generate human disease models for research, various mouse models, including knockout, knockin, humanization, point mutation, and transgenic models, have been produced with the assistance of embryo manipulation methods. Endogenous mouse genes can be modified or exogenous DNA may be inserted in the mouse genome through various manipulation tools. These tools include somatic cell nuclear transfer, microinjection of exogenous DNA into the pronucleus of a fertilized oocyte, microinjection of genetically modified stem cells into a blastocyst, and intracytoplasmic sperm injection (ICSI). The first genetically engineered animal was produced by injecting the Simian Virus 40 (SV40) DNA into mouse embryos in 1974, which ushered in a new era of transgenic animal models [3].
The emergence of programmable nucleases, such as zinc finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs), and clustered regularly interspaced short palindromic repeats (CRISPR)/Cas, has facilitated the production of genetically modified mouse (GEM) models with precise site-specific targeting abilities [4]. These programmable nucleases generate DNA double-strand breaks (DSBs) at the target sequences, which can be modified or repaired using DNA repair pathways, homology-directed repair (HDR), or non-homologous end joining (NHEJ). These repair processes result in insertion/deletion (indel) mutations, insertion of particular sequences, and correction of disease-associated variants at the target sequences. Recently, base editing (BE) and prime editing (PE) have been successfully developed and utilized to introduce more diverse mutations into the target sites [5,6]. When mouse models are generated using programmable nucleases, the conventional methods used are cytoplasmic or pronuclear microinjection or electroporation.
However, despite the ample studies describing the applications of engineered nucleases for developing mouse models of human diseases, comprehensive knowledge of the best practices for effective generation of mouse models as well as therapeutics based on nucleases need to be improved further. In the present review, we summarize the current progress in and prospects for developing mouse models using programmable nuclease-mediated gene editing technologies and their therapeutic applications.

2. Programmable Nucleases

Since the early-phase studies on genome editing, ZFNs and TALENs have been used to generate DSBs at the DNA target sites [7,8], and their structural features and editing mechanisms have been reviewed extensively [7,8,9,10]. Briefly, ZFN consists of a nonspecific cleavage domain of the Flavobacterium okeanokoites-derived FokI and a DNA-binding zinc-finger domain. About 3–6 tandem Cys2-His2 zinc fingers constitute a zinc-finger domain, each of which binds the corresponding three nucleotides [7,9,10]. Because of this hybrid structure, ZFNs are versatile and specific for DNA binding and robust cleavage. Therefore, ZFNs have emerged as potent tools for genome editing [9,10]. Similar to ZFNs, TALENs also exploit the FokI endonucleases for DNA cleavage. However, the DNA-binding domains in TALENs are transcription activator-like effector (TALE) repeats derived from the plant pathogenic Xanthomonas bacteria [8,11]. Compared to ZFNs, the construct of TALENs is less complex, and their application as molecular scissors is more specific and practical [12]. Even with this versatility and robustness, ZFNs and TALENs have several limitations, such as the need for reassembly of the ZFP or TALE modules for each target sequence. In addition, the activity of ZFNs is greatly dependent on the interaction specificity of adjacent fingers, which makes their overall function unpredictable, impeding their broad application [4,7,9,10]. Compared to ZFNs and TALENs, the CRISPR/Cas system containing a Cas nuclease and a single-guide RNA (sgRNA) can be used to modify target DNA sequences in a simpler manner [13,14]. Although this system may cause unwanted off-target effects, these can be minimized by selecting the sgRNA to have more specificity for the target sequence [15,16]. Of the Cas nucleases, the Cas9 nuclease has been commonly used as a genome editing tool, whose protospacer adjacent motif (PAM) sequence is 5′-NGG-3′. Another nuclease, Cpf1, known as Cas12a, uses a T-rich PAM, such as 5′-TTTV-3′, and exhibits various advantages over the Cas9 nuclease, including more frequent PAM sequences in the genome, smaller nuclease size, and lower incidence of off-target effects. Therefore, Cpf1 is an alternative to Cas9 nuclease and is beneficial for introducing knockin mutations when PAM sequences near the target site are less efficient [17]. Among the Cpf1 orthologs, AsCpf1, FnCpf1, and LbCpf1 nucleases have been frequently and widely utilized in gene editing [18,19,20,21].
CRISPR/Cas13, a recently developed RNA guide platform, recognizes single-stranded RNA instead of the double-stranded DNA [17,22,23,24,25]. The Cas13 enzyme has two higher eukaryote and prokaryote nucleotide-binding nuclease domains (HEPN), which degrade both the target and non-target RNAs randomly [22,23]. Mainly, Cas13 nuclease is more versatile than Cas9 or Cpf1, because it does not require PAM at the target site. In addition, RfxCas13d is the smallest protein that shows the most effective activity for RNA knockdown, regulation of alternative splicing, and viral resistance among the Cas13 orthologs [23,26,27,28,29].
Because of this facilitation and high applicability, genome editing tools have great potential for correcting mutations in human genetic disorders. However, their applications are limited owing to the generation of DSBs, which can cause indiscriminate indels at the target sites. Recently, the discovery of BE and PE has opened new opportunities to alleviate this limitation and hurdle because they do not generate DSBs or require donor DNA for gene corrections [30]. These editors have been demonstrated to be more efficient with lower off-target effects in vitro and in vivo than CRISPR/Cas9 [30,31]. Two BEs, cytosine BE (CBE) and adenine BE (ABE), have been reported; CBE consists of a nickase Cas9 (nCas9, D10A) or a dead Cas9 (dCas9, D10A, and H840A), a cytidine deaminase, and an uracil glycosylase inhibitor; ABE consists of nCas9 or dCas9 and evolved deoxyadenosine deaminases [31,32,33]. CBE replaces cytosine (C) with thymine (T) and ABE replaces adenine (A) with guanine (G) at the target site [31,32,33]. PE comprises a nickase Cas9 and a reverse transcriptase (RT); it has been known to be more versatile than CRISPR/Cas9, owing to its ability to generate various gene editing variants, including base transfers and small indels [6,31,33,34]. The recently developed BE and PE have been reported to have higher efficiencies and expanded target scopes in gene corrections [30,31].
Overall, owing to the straightforward, effective, and precise characteristics of the Cas-based systems, they are more extensively utilized for gene modifications [13,35]. These gene editing tools offer the opportunity to revolutionize the generation of mouse models with disease-associated variants.

3. Generation of Mouse Models of Human Diseases Using Genome Editing Tools

3.1. Mouse Models with Gene-Knockout or Point Mutations

Knockout strategies in mice using programmable nucleases have been reported in several studies [36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52]. Direct injection of the ZFN platforms into mouse embryos result in inactivation of the target genes, such as Notch3, Mdr1a, and Jag1, to generate GEM models [36]. Although ZFN-based methods have been effectively used to generate knockout mice [36,37,43], the complexity of target-specific ZFN manipulation and zygote toxicity have limited their extensive applications [37] (See Table 1 for Summary).
The first reports of TALEN-mediated knockout mice involved Pibf1 and Sepw1 knockout mice that were generated by injecting TALEN mRNA into mouse embryos [38], in which about 49–77% of the pups had the mutation. In a subsequent study, the microinjection of TALEN mRNAs into mouse zygotes resulted in the knockout of several miRNAs, such as mmu-mir-10a and -10b [39]. In addition, another study found that applying two TALEN pairs targeting the FATS functional domain of the CFS gene resulted in larger deletions in size and bi-allelic mutations as well as further increase in the mutation efficacy compared to a single TALEN pair [40]. The cytoplasmic microinjection of Ttc36-targeted TALEN improved the knockout efficiency approximately three-fold compared to its injection into the pronuclei of mice [41]. The microinjection of TALEN mRNAs into the oocytes from CD1, C3H, and C57BL/6J mouse strains efficiently produced Zic2 knockout mice [45]. These findings indicate that TALEN-mediated gene targeting is more effective in generating knockout mice than ZFNs because of the enhanced survival of the pups and higher efficiencies of introducing mutations.
It is well known that the CRISPR/Cas9 system is a powerful tool for producing GEMs in a highly efficient, simple, versatile, and cost-effective manner. Since the advent of the CRISPR/Cas9-mediated mutant mice, this approach has wholly replaced several platforms for genome manipulations that have been used commonly for over 30 years. CRISPR/Cas9-mediated enhanced green fluorescent protein (EGFP) knockout was achieved by co-injecting Cas9 mRNA and EGFP sgRNA into transgenic mice [53]. In addition, the cytoplasmic injection of Tet1- and Tet2-targeted Cas9 mRNA/sgRNA into mouse zygotes has been demonstrated as a reliable and efficient method for generating bi-allelic mutations [42]. The cytoplasmic microinjection of Cas9 mRNA and sgRNA has been reported as the most efficient method for gene knockout in mice as opposed to pronuclear microinjection [52,54]. The microinjection of Cas9 mRNA and sgRNA into the pronuclei of mice following in vitro fertilization also generated indel mutations in the Fah allele in NOD-Rag1nullIL2Rγnull mice [55]. The cytoplasmic injection of Cas9 and truncated sgRNA complexes were reported to disrupt FVII with high efficiency compared to standard complexes composed of Cas9 and standard sgRNA [44]. CRISPR/Cas9 has been utilized to disrupt the ATP6V1H gene, which causes bone loss in mice [56]. Deleting the Notch3 gene via the injection of Cas9/sgRNA into the mice embryos was reported to generate a human lateral meningocele syndrome-associated mouse model, which displayed a reduction in several femur trabecular scores, such as the cancellous bone volume, connectivity, and trabecular number [57]. Pankowicz et al. reported a knockout mouse for the liver, in which the CRISPR/Cas9 system deleted both Fah and Hpd or Hpd and Asl genes. These mice could be used to develop liver-disease-related models as well as further investigate metabolic liver disorders and liver function-related genes [58,59]. A previous study rapidly and effectively generated a mouse model with a 5 Mb deletion of the SOX9 gene by microinjection of Cas9/sgRNAs and a donor vector into fertilized eggs [60]. Similarly, various studies have reported the deletions of multiple genes, such as the 0.5 Mb (Irx3 and Irx5 genes), 1.15 Mb (intervals from Nox4 to Grm5 genes, including Tyr locus), 1.2 Mb (regions carrying nine genes, including Srd5a3, Exoc1, Nmu, Kit, Kdr, Clock, Cep135, Tmeme165, and Pdcl2), 30 kb (Pcdh), and 95 kb (Fpr1-3). However, the efficiencies of these large deletions were relatively low [51,61,62,63,64]. A previous study reported acampomelic campomelic dysplasia (ACD) and campomelic dysplasia (CD) mouse models in which the CRISPR/Cas9 system was used to knock out the SOX9 gene, a transcription factor involved in the development of cartilage [65]. Furthermore, adeno-associated viral (AAV)-expressing sgRNA for p53, LKB1, and KRAS genes was delivered to the lung of mice expressing CRISPR-Cas9 via tail vein injection, resulting in target gene mutations and lung tumorigenesis [66].
Since the development of base editing technologies capable of editing specific nucleotides, the CBE method has been used to generate Tyr or Dmd mutants with C to T conversion by delivering sgRNA and CBE mRNA or ribonucleoprotein (RNP) to mouse zygotes [67]. Subsequently, the CBE-based CRISPR-stop method has been used to generate mutant mouse models with stop codons by converting single nucleotides in the coding sequences of various genes, such as Tyr, vGlut3, Slc26a5, Otof, and Atoh1 [68]. Wang et al. reported the generation of a PD-1 mutant mouse using a novel strategy known as induce gene silencing (i-Silence), with an ABE-mediated start codon mutation from ATG to GTG or ACG [69]. Using the PE system, a mouse model has been generated to have one deletion in exon 3 of the Crygc gene by microinjection of the PE3 plasmid directly into the mouse zygote that shows a cataract phenotype [70]. These findings provide a promising foundation that such editing tools can be extensively used for editing or correcting human disease-associated mutations.
Table 1. Generation of mouse models using gene editing tools. NA: not available, without information of related diseases in general.
Table 1. Generation of mouse models using gene editing tools. NA: not available, without information of related diseases in general.
Gene Editing ToolTarget GeneHuman Disease ModelsMutation TypeManipulation MethodsReference
ZFNNotch3, Mdr1a, and Jag1NAKnockoutMicroinjection[36]
TALENPibf1 and Sepw1NAKnockoutMicroinjection[38]
TALENmmu-mir-10a and -10bNAKnockoutMicroinjection[39]
TALENTtc36NAKnockoutMicroinjection[41]
TALENZic2NAKnockoutMicroinjection[45]
CRISPR/Cas9Tet1 and Tet2NAKnockoutMicroinjection[42]
CRISPR/Cas9FahTyrosinaemiaKnockoutMicroinjection[55]
CRISPR/Cas9FVIIHemophiliaKnockoutMicroinjection[44]
CRISPR/Cas9ATP6V1HOsteoporosis/
osteopenia
KnockoutMicroinjection[56]
CRISPR/Cas9Notch3Lateral meningocele syndromeKnockoutMicroinjection[57]
CRISPR/Cas9Fah and HpdTyrosinaemia type IKnockoutHydrodynamic injection[58]
CRISPR/Cas9Hpd and AslHyperammonemiaKnockoutHydrodynamic injection[59]
CRISPR/Cas9SOX9Acampomelic campomelic dysplasia (ACD)KnockoutMicroinjection[65]
CRISPR/Cas9p53, LKB1, and KRASLung adenocarcinomaKnockoutMicroinjection[66]
CBEsDmdNAKnockoutMicroinjection/Electroporation[67]
CBEsTyrPigmentation disorderKnockoutMicroinjection/Electroporation[67]
CBEsAtoh1, vGlut3, Slc26a5, OtofHearing lossKnockoutMicroinjection[68]
ABEsPD-1NAKnockoutMicroinjection[69]
PECrygcCataract disorderKnockoutMicroinjection[70]
ZFNRab38NAKnockinMicroinjection[71]
TALENRab38NAKnockinMicroinjection[72]
TALENFusAmyotrophic lateral sclerosisKnockinMicroinjection[73]
TALENCrb1Retinal diseaseKnockinMicroinjection[74]
TALENScn8aEpileptic encephalopathyKnockinMicroinjection[75]
TALENOct4 and NanogNAKnockinMicroinjection[76]
CRISPR/Cas9ActbNAKnockinMicroinjection[77]
CRISPR/Cas9BRILOsteogenesis imperfecta (OI) type VKnockinMicroinjection[78]
CRISPR/Cas9HTTHuntington’s disease (HD)KnockinMicroinjection[79]
ZFNCCR5HIV-1HumanizedTransplantation/Engraftment[80,81,82]
TALENHBGβ-thalassemia and sickle cell diseaseHumanizedTransplantation/Engraftment[83]
CRISPR/Cas9DMDDMDHumanizedMicroinjection[84]
CRISPR/Cas9Pdcd1 and Lag3NAHumanizedElectroporation[85]

3.2. Knockin Mouse Models

Over the last ten years, the number of newly developed mouse models has increased dramatically, which has extended the applications of programmable nucleases. The first report of ZFNs for the generation of knockin models was for the insertion of the GFP-linked β-Galactosidase gene into the Rosa26 locus with a 1.7–4.5% targeting efficiency [43]. Other knockin models have also been produced by inserting reporter cassettes into the Rosa26 or Mdr1a locus using ZFNs [37,86]. It was reported that the injection of Rab38-targeted ZFNs or TALENs with a 144-base single-stranded oligodeoxynucleotide (ODN) into single-cell embryos resulted in homologous recombination (HR)-mediated modifications with approximately 2% efficiency [71,72]. In addition, the fusion of TALEN-95As, an improved plasmid-coded poly(A) targeting the Fus gene, and the 140-base mutant ODN has been used to produce an amyotrophic lateral sclerosis (ALS)-related mouse knockin model with increased knockin efficiency of up to 6.8% [73]. Microinjecting TALEN mRNAs into the pronucleus generated TALEN-mediated knockin mice for the target genes, such as Satb1, Crb1, and Scn8a [74,75,87]. The injection of EGFP/mCherry reporter-linked Oct4 and Nanog vectors, Cas9 mRNA, and sgRNAs into either the cytoplasm or pronucleus resulted in effective insertion with 9–19% efficiency [76]. In addition, an injection experiment with a mixture containing an EGFP-linked Actb vector, Cas9 protein, Actb crRNA, and tracrRNA into the pronuclei showed increased insertion efficiency by 50%. These results suggest that this is one of the practical and convenient strategies for producing knockin mouse models [77]. Another knockin mouse model using CRISPR/Cas9 has been generated to induce BRIL variants found in patients with osteogenesis imperfecta type V [78].
The CAG repeat sequences in huntingtin (HTT) gene have been reported to cause Huntington’s disease (HD) [88]. Yang et al. generated a knockin HD mouse model with CAG repeat sequences in the HTT genes using CRISPR/Cas9. They evaluated the toxicity of the repeat-associated non-AUG-mediated products in the mouse model [79]. This development of HTT knockin mice using CRISPR/Cas9 is expected to be a promising breakthrough for researching pathogenesis and therapeutics for HD disease.

3.3. Humanized Mouse Models

Humanized models can be generated by transplanting short strands of DNA sequences, cells, tissues, or microbiomes from humans to mice. They are broadly exploited in the study of various human diseases, such as HD, human immunodeficiency virus type 1 (HIV-1), Charcot–Marie–Tooth type 1 (CMT1), and cancers [80,89,90,91]. The generation of humanized mouse models using gene editing technologies typically involve insertion of human-derived DNA sequences at specific sites or transplanting of gene-edited human-derived cells into mice.
Using CCR5-targeted ZFNs, the CCR5 gene, a co-receptor playing a fundamental role in HIV-1, was disrupted in human hematopoietic stem/precursor cell (HSC) with a mutant efficiency of 17% [80,81,82]. Edited HSCs were transplanted in NOD/SCID/IL2Rγnull (NSG) mice, which can be a potential model for developing HIV-1 treatment because the deletion in the HSCs is maintained even after maturation [80,81,82]. β-thalassemia and sickle-cell disease are hemoglobin disorders caused by pathogenic mutations of the HBG gene. TALENs targeting the HBG promoter were employed to drive the hereditary persistence of the fetal hemoglobin phenotype (6%) in human CD34+ cells. The edited cells were transplanted in W41 mice and were investigated for persistence in hematopoietic cells for 24 weeks [83]. Duchenne muscular dystrophy (DMD) is a muscular disorder caused by mutations in the dystrophin gene on the X chromosome [92]. A Δ50;h51KI humanized DMD mouse model was generated using CRISPR/Cas9 targeting the DMD gene; in these mice, exon 51 was replaced with human exon 51 and exon 50 was deleted [84]. To generate human PD-1/LAG-3 knockin mice expressing human PD-1 protein, mouse Pdcd1 and Lag3 genes were replaced with human Pdcd1 and Lag3 genes using CRISPR/Cas9 [85]. This dual-immune checkpoint model is expected to enable evaluation of human immune responses, including the analyses of cytokine release, tumor growth, and lymphocyte populations.

3.4. Transgenic Mouse Models

Transgenic mice can be generated by microinjecting exogenous recombinant DNA directly into the pronuclei of fertilized oocytes. The recombinant DNA is randomly inserted in the chromosome and expressed by an exogenous promoter [93]. However, the practical application of this method is limited as the expression of the transgene is critically affected by the integration site in the chromosome [94].
Recently, CRISPR/Cas9-based conditional transgenic mice harboring Rosa26-LSL-dCas9-p300 or Rosa26-LSL-dCas9-KRAB have been generated, which express transgenes in a Cre-dependent manner. The Rosa26-LSL-dCas9-p300 mice were designed for gene activation and Rosa26-LSL-dCas9-KRAB for gene repression. The target genes were regulated by targeting gRNAs to the transcriptional start sites or distal enhancer elements in the transgenic mice, resulting in corresponding changes in the epigenetic states and downstream phenotypes in the brain and liver [95]. These transgenic mice could be applied to more diverse functional and mechanism studies using tissue-specific Cre and sgRNAs.

4. Therapeutic Applications Using Genome Editing Tools

4.1. Duchenne Muscular Dystrophy (DMD)

DMD, an X-linked recessive disorder, is caused by a mutation in the DMD gene, which impairs normal dystrophin function. The mouse model with a point mutation in the DMD gene is the mdx mouse, which is the most popular model for studying DMD [96,97,98,99,100]. Several recent studies have attempted to restore dystrophin function by eliminating DMD mutations using programmable nucleases [96,97,98,99,100]. Compared to ZFNs and TALENs, the CRISPR/Cas9 system has been applied more commonly in studies on treating DMD. Several studies have reported the AAV-mediated CRISPR/Cas9 platform for correcting DMD mutations in dystrophic mdx mice and observed significantly improved muscular functions in the treated mice [96,97,98,99,100]. In addition, CRISPR/Cas9 has been utilized to generate a novel DMD mouse model with the disruption of exon 50 of the DMD gene, which exhibits severe muscle dysfunction. Treatment with AAV9 expressing both Cas9 and exon-51-targeted sgRNA in the mouse resulted in restoration of dystrophin expressions in the cardiac and skeletal muscles by up to 90% [101]. The RNA-guided endonuclease CRISPR/Cpf1 has been used to correct a nonsense mutation in exon 23 in mdx mice, which was efficiently and permanently restored by injecting LbCpf1 mRNA, 180-base ssODN, and sgRNA into the mouse zygotes [19]. Various follow-up studies have reported that AAV9-mediated CRISPR/Cas9 delivery effectively restored dystrophin expression in DMD mouse models harboring deletions of exons 43, 44, 45, 51, or 52 [84,102,103,104]. Another study showed that these gene corrections and dystrophin protein expressions by AAV-CRISPR treatment were maintained in mdx mice for up to one year [105]. Although the development of AAV-CRISPR as a therapeutic method must consider unwanted gene editing and transcriptome modifications, it is undeniable that AAV-CRISPR may offer the potential to correct the genome permanently [105] (See Table 2 for Summary).
Ryu et al. reported that a DMD-gene-targeted ABE7.10 base editor was delivered into mouse muscle using double trans-splicing AAV and that up to 17% of exon 20 mutations in the DMD mice were corrected without off-target effects [106]. However, the large size of ABE7.10 with sgRNA (~6.1 kb) limits the common use of these BEs [106]. To address this limitation, iABE-NGA, a mini adenine BE targeting NG, was developed. Delivery of the AAV-packaged iABE-NGA into mdx4cv mice precisely and effectively corrected DMD mutations and recovered dystrophin expression without toxicity and with low off-target effects [107]. In addition, the delivery of CRISPR/Cas9 RNP complexes by nanoparticles, such as 5A2-DOT-X LNPs, into the tibialis anterior muscles of the exon-44-deleted DMD mice restored dystrophin expression by 4.2% [108]. Recently, CRISPR/Cas9 was used to correct induced myogenic progenitor cells (iMPCs) derived from DMDmdx mice with mutations in exon 23. The corrected iMPCs were subsequently engrafted into dystrophic limb muscles to successfully restore dystrophin protein expression [109]. Overall, these approaches for gene correction in DMD models might provide promising strategies for repairing pathogenic mutations in patients with muscular dystrophies.
Table 2. Gene editing-mediated therapeutics using mouse models of human disease.
Table 2. Gene editing-mediated therapeutics using mouse models of human disease.
Human DiseaseMutant GeneMutation TypeGene Editing ToolDelivery
Method
Reference
Duchenne muscular
dystrophy (DMD)
DMDKnockoutCRISPR/Cas9, CRISPR/
Cpf1, and ABE
AAV/RNP[105,106,107,108,109]
Hemophilia AF8KnockoutCRISPR/Cas9AAV[110,111,112]
Hemophilia BF9KnockoutZFN and CRISPR/Cas9AAV[113,114,115,116,117,118]
Huntington’s
disease (HD)
HTTHumanizedZFNs, CRISPR/Cas9,
CRISPR/Cpf1,
and RfxCas13d
AAV[18,90,119,120,121,122,123]
Amyotrophic lateral
sclerosis (ALS)
SOD1TransgenicCRISPR/Cas9,
RfxCas13d, and CBE
AAV[124,125,126,127]
Hearing lossTMC1/
Kcnq4
KnockoutCRISPR/Cas9,
RfxCas13d, PE, and CBE
AAV/RNP[128,129,130,131,132,133]
Hereditary tyrosinaemia
type I (HT-I)
FahKnockoutCRISPR/Cas9, ABE, and PEAAV/LNP/Plasmid[58,134,135,136,137,138,139,140]
Neovascular age-related macular degeneration
(AMD)
Vegfa/
HIF-1α
KnockoutCRISPR/Cas9
and LbCpf1
AAV/RNP/
Lentivirus
[141,142,143,144]
Charcot-Marie-Tooth
disease (CMT1)
PMP22HumanizedCRISPR/Cas9RNP[89,145]
Leber congenital
amaurosis (LCA)
Rpe65KnockoutCRISPR/Cas9, ABE, and PEAAV/Lentivirus/
eVLPs
[146,147,148,149]
Alzheimer’s disease
(AD)
Bace1/
Mt1
TransgenicCRISPR/Cas9RNP/Lentivirus[150,151]
Familial
hypercholesterolemia
(FH)
LdlrKnockoutCRISPR/Cas9AAV[152]
Pompe disease (PD)GaaKnockoutCRISPR/Cas9Electroporation[153,154]
ProgeriaLmnaKnockoutCRISPR/Cas9AAV[155,156]

4.2. Hemophilia

Hemophilia is an X-linked hemorrhagic disorder characterized by the deficiency of clotting factors VIII (F8, Hemophilia A) or IX (F9, Hemophilia B) [157]. Injecting the Alb intron-13-targeted AAV8-SaCas9-gRNA and AAV8-modified human B domain deleted-F8 into FVIII hemophilic knockout mice resulted in elevation of their plasma FVIII levels and amelioration of blood loss without liver toxicity and off-target effects [110]. Other studies have reported the rescue of F8 expression in hemophilia A using CRISPR/Cas9 and AAV delivery [111,112]. Anguela et al. found that the injection of AAV-ZFN expressing hF9 gene and AAV-donor harboring homologous arms to the target gene into hF9 mutant mice resulted in an approximately five-fold increase in the expression of hF9 protein compared to a previous study [113,114]. A hemophilia B mouse model carrying the Y371D mutation in the F9 gene was previously generated using the CRISPR/Cas9 system [115]. Mutant mice were used for gene correction experiments in which Cas9 and F9-targeted sgRNA components were delivered via adenoviral (AdV) or naked DNA vectors. Direct injection of the naked DNA resulted in correction of more than 0.56% of the F9 mutations, which are sufficient for restoring the hemostasis phenotype. In contrast, the AdV delivery showed no therapeutic application because of severe toxicity in the hepatocytes of the mice, despite the remarkable repair efficacy [115]. Another study demonstrated that the AAV-mediated delivery of CRISPR/Cas9 to the hemophilia B mouse model recovered FIX expression [116,117,118]. Recently, the injection of dual AAV-CjCas9 targeting hF9 and a donor template into the hemophilia mice induced hFIX expressions and improved coagulation functions [116]. Together, these experimental data strongly support the observation that genome editing tools can be used therapeutically in the clinical treatment of hemophilia.

4.3. Huntington’s Disease

Huntington’s disease (HD), one of the neurodegenerative disorders, is characterized by the expansion of CAG repeats that encode polyglutamine in the HTT gene [88]. Isalan et al. reported that the delivery of a striatal AAV vector harboring zinc fingers (ZFs) targeting pathogenic CAG-repeat expansions into the R6/2 mice, an HD mouse model, resulted in the suppression of cerebral HTT expression and reduction of HD-associated symptoms [119]. In addition to ZFs, AAV-mediated deliveries of CRISPR/Cas9 and CRISPR/Cpf1 have been ubiquitously utilized to remedy HD via HTT gene editing [18,120,121,122,123]. Delivery of AAV1 packaging RfxCas13d, a Cas13 variant targeting hHTT exon 1 in R6/2 HD mice, decreased HTT expression by approximately 50%, suggesting the ability of RfxCas13d to ameliorate HD-linked pathology [124]. This finding indicates that RfxCas13d may be a potential tool for suppressing HTT and neurodegeneration-related genes in the central nervous system.

4.4. Amyotrophic Lateral Sclerosis

Amyotrophic lateral sclerosis (ALS) is a neurodegenerative disorder characterized by the malfunction of nerve cells in the brain [158]. SOD1 mutations are the most prevalent and fundamental cause of ALS [159]. Because CRISPR/Cas effectively interferes with the expressions of the mutant genes in SOD1 ALS mouse models, this editing technology has been considered a potential therapeutic tool for ALS [124,125,126,127]. Duan et al. reported that CRISPR/Cas9 targeting SOD1G93A mutations was delivered by AAV9 vehicles to ALS transgenic mice, disrupting the mutant SOD1 gene and increasing the lifetime of the mice by up to 54.6%. Furthermore, the ALS phenotype of the mice, including muscle atrophy and paralysis, was improved [126]. The AAV9-packaged RfxCas13d targeting mutant SOD1 reduced 50% of the expression of SOD1 in the ALS spinal cord, followed by functional improvement of neuromuscular disorders as well as decrease in muscle atrophy [124]. Interestingly, dual-AAVs-delivered split-intein cytidine base editors reduced the expression of mutant SOD1, leading to the amelioration of muscle denervation and muscle atrophy and improvement of neuromuscular function in ALS mice [127]. Together, these CRISPR-mediated therapeutic applications for ALS have great therapeutic potential because they have become more sophisticated and diverse.

4.5. Hearing Loss

Hearing loss is one of the most prominent disorders among newborns and is caused mainly by mutations in the protein transmembrane channel-like 1 (Tmc1) gene [160]. Several studies have reported gene therapy strategies targeting Tmc1 using CRISPR/Cas9 or CBE [128,129,130]. Mutant Tmc1-targeting CRISPR/Rfx Cas13d was packaged in AAV-PHP.eB, a previously developed novel AAV vector system [131]. Gene editing with this CRISPR/Cas system reduced the expression of Tmc1 by up to 70.2% in the cochlea of Bth neonatal mice, resulting in alleviation of the pathological symptoms in the mice [132].
Degeneration of the outer hair cells (OHCs) has also been reported to cause hearing loss [161]. Noh et al. generated a Kcnq4 mutant mouse exhibiting an OHC degenerative phenotype. When dual AAV vectors or RNP harboring SpCas9 and sgRNA were injected, the Kcnq4 mutant mouse showed partial recovery of hearing function [133]. These results are strong evidence that the therapeutic applications of CRISPR/Cas9 are promising alternatives for treating hearing loss.

4.6. Hereditary Tyrosinemia Type I

Hereditary tyrosinemia type I (HT-I) is an inherited metabolic disorder that results from a mutation in the Fah gene encoding fumarylacetoacetate hydrolase [162]. When Cas9 and Fah-targeting sgRNA were injected into Fah5981SB mice carrying a homozygous G to A variant located at the last base of exon 8 of the gene, the pathological phenotypes of the mice were improved [134]. VanLith et al. transduced a lentiviral vector harboring Cas9 nuclease and Fah locus-targeting sgRNA as well as an AAV particle carrying a 1.2 kb homologous template to the Fah−/− hepatocytes. When these hepatocytes were transplanted in Fah−/− mice, they found that the edited hepatocytes improved liver injuries and restored metabolic disorder [135]. Since then, several studies have demonstrated that editing the mutant Fah gene via the delivery of AAV-encoding CRISPR or PE components could help recover the lethal phenotype of the Fah-deficient mice [58,136,137]. Although successful recoveries of the pathogenic phenotypes were observed, the AAV system has critical limitations for common application. The AAV vector can allow DNA fragments of up to 4.7 kb because of its loading capacity. Recently, a promising alternative method has been reported: delivery of CRISPR/Cas9 components, including ABE mRNA and Fah-targeting sgRNA by lipid nanoparticles (LNPs) [138,139]. Jiang et al. reported that a PE-Cas9-based deletion and repair (known as PEDAR) method could recover the expression of functional FAH in a Tyrosinemia I mouse model by accurately eliminating a 1.38 kb pathogenic fragment of the Fah gene via liver-targeted hydrodynamic injection [140]. These results demonstrate the outstanding application of PEDAR in the repair of Fah in Tyrosinemia I mice and its potential for correcting pathogenic genes caused by insertion of large fragments.

4.7. Age-Related Macular Degeneration (AMD)

Neovascular AMD is one of the most prevalent causes of vision loss globally [163]. Several studies have adapted gene editing tools to modulate the vascular endothelial growth factor (Vegfa) and hypoxia-inducible factor-1α (HIF-1α) associated with AMD [141,142,143,144]. Co-delivery of two separate lentivirus particles harboring Cas9 mRNA and sgRNA to an AMD mouse model resulted in 44% deletion of Vegfa in the retinal pigment epithelial (RPE) cells and 63% decrease in choroidal neovascularization (CNV). Kim et al. exploited the subretinal injection of RNPs, composed of Cas9 and sgRNA, into an AMD mouse model to cause mutations in Vegfa [144]. Another study reported that gene editing with Cas9 RNPs successfully reduced CNV in AMD mice [143]. When the AAV-packaged LbCpf1 targeting Vegfa or Hif1a was intravitreally injected into laser-treated mouse retinas, CNV was effectively reduced without detectable cone damage. It has been reported that the LbCpf1 approach is more effective than Aflibercept, a VEGF inhibitor commonly used in clinical practice, for CNV reduction. Moreover, these effects of the LbCpf1 system lasted longer without repeated injections [142]. Collectively, these results suggest that gene editing tools with LbCpf1 could be potent therapeutic agents for treating AMD.

4.8. Charcot–Marie–Tooth Type 1

Charcot–Marie–Tooth type 1 (CMT1) is a disease involving various genetic abnormalities that characterize the phenotypes of motor and sensory neuropathies [164]. The duplication of peripheral myelin protein (PMP22) explains 70% of CMT1 cases [164]. Recently, CRISPR/Cas9 RNA complexes targeting the TATA-box of hPMP22 were intraneurally injected into a C22 mouse, resulting in a decrease in PMP22 expression and reduction of demyelination [145]. This is a pioneering report indicating that the CRISPR/Cas9 system could be a promising therapeutic strategy for treating CMT1 disease.

4.9. Leber Congenital Amaurosis

Leber congenital amaurosis (LCA) is one of the leading causes of blindness in infants. Among LCA-linked genes, RPE65 mutations account for approximately 6% of the LCA cases [165]. In a previous study, two separate AAV vectors expressing Cas9 and including a donor template linked to a TS4 sgRNA targeting a mutation in exon 3 of the Rpe65 gene were delivered to rd12 mice, a naturally occurring model of LCA with a Rpe65 mutation, via subretinal injection. The treated LCA mice showed improved pathogenic phenotype by deleting approximately 1.6% and causing more than 1% HDR of the Rpe65 mutation [146]. Subretinal injection of ABE- and sgRNA-expressing lentiviral particles in the rd12 mice efficiently corrected mutations in the Rpe65 gene (29%) with low indels and no off-target effects. ABE treatment in the Rpe65 mutant mice increased RPE65 expressions and restored visual functions, which were mediated by functional restoration and prolongation of survival of the cone photoreceptors [147,148]. Jang et al. recently generated an AAV vector encoding Rpe65-targeting pegRNA and PE2; subretinal injection of the vector into rd12 mice resulted in efficient editing of the target gene by up to 6.4% without any unwanted editing and revealed dramatic improvements to visual function [149]. Together, these results indicate that high-precision editing using PE2 and AAV delivery may be a potential option for treating patients with LCA.

4.10. Alzheimer’s Disease

Alzheimer’s disease (AD) is the most common neurodegenerative disorder characterized by slow memory destruction, synaptic impairment, and Aβ accumulation [166]. Since the first report of type 1 melatonin receptor (Mt1) and β-secretase 1 (Bace1) for the regulation of Aβ production, targeting Mt1 and Bace1 has been considered a powerful therapeutic approach for Alzheimer’s disease [167]. Park et al. generated amphiphilic nanocomplexes loaded with Cas9 and Bace1-targeting sgRNA and then injected into the hippocampi of 5XFAD and APP transgenic AD mouse models. This approach effectively reduced Bace1 expression and cognitive deficiencies to alleviate the Aβ-associated pathologies in the AD mouse models [150]. In addition, injection of an Mt1-Cas9 activator into the brain of the 5xFAD mice efficiently activated Mt1 expression and relieved cognitive deficits in AD mice [151].

4.11. Familial Hypercholesterolemia

Familial hypercholesterolemia (FH) is an inherited disease characterized by significant elevation of the low-density lipoprotein cholesterol [168]. This disease is mainly caused by mutations in the Ldlr gene [169]. Zhao et al. developed CRISPR/Cas9-loaded AAV particles to correct these Ldlr mutations; the particles were transferred to the hepatocytes of a LdlrE208X mouse model, which effectively reduced atherosclerosis phenotypes in the mutant mice [152]. Overall, this finding demonstrates that Ldlr editing using AAV-CRISPR/Cas9 can serve as a potential therapeutic strategy for treating FH.

4.12. Pompe Disease

Pompe disease (PD) is an autosomal recessive disease caused by mutations in the GAA gene, which encodes acid α-glucosidase, an essential enzyme that hydrolyzes lysosomal glycogen [170]. Through microinjection into fertilized zygotes using CRISPR/Cas9-mediated HDR, mice with infantile-onset PD were generated expressing Gaac.1826dupA or GaaEm1935C>A. The mice exhibited hypertrophic cardiomyopathy and skeletal muscle deficiency associated with PD [153,154]. This PD mouse model can be used for gene therapy research targeting Gaa mutations as well as mechanisms related to PD.

4.13. Progeria

Progeria, also known as Hutchinson–Gilford progeria syndrome (HGPS), is a rare genetic disorder characterized by premature aging in children, leading to early death [171]. De novo point variants in the LMNA gene are known to cause 90% of HGPS cases [172]. Fernández et al. and Beyret et al. delivered an AAV9-packaged CRISPR/Cas9 system targeting LMNA exon 11 intraperitoneal injection into LmnaG609G/G609G HGPS mice. Despite the low editing efficacy of this approach, it not only improved the HGPS-specific phenotypes but also extended the lifespans of the HGPS mice [155,156]. Although further studies are needed to evaluate the side effects and off-target effects of CRISPR/Cas9, this finding opens up the possibility of applying CRISPR/Cas9 to HGPS treatment.

5. Conclusion and Future Prospects

Recent technological advances in programmable nucleases, including ZFNs, TALENs, and CRISPR/Cas, have significantly contributed to the development of therapeutic strategies for various diseases in human. In particular, ZFNs, TALENs, and CRISPR/Cas9 have been applied in clinical trials [173,174]; among these three genome editing tools, CRISPR/Cas has been proven to regulate the expressions of target genes simply and shows promise for development into various versatile tools. Accordingly, novel miniature CRISPR/Cas12f nucleases, such as AsCas12f1, Un1Cas12f1, and SpaCas12f1 (422, 529, and 497 amino acids, respectively), have recently been indicated as effective editing tools in bacteria, cells, and tissues using delivery strategies, such as plasmids, RNPs, and AAV [175,176,177]. CRISPR-Casπ (Cas12I) is a recently developed miniature type V CRISPR/Cas system comprising 860 amino acids that has demonstrated significant editing capacity for cells compared to two well-engineered editors, namely LbCas12a and SpyCas9 [178]. A recent study reported Cas12f-based ABEs and CBEs with engineered TadA, which show successful gene editing with minimal off-target effects both in vitro and in vivo via the all-in-one AAV delivery system [179]. These mini gene editors can be easily packaged in an AAV vehicle to provide promising therapeutic options for various human diseases (See Figure 1 for Summary).
Although the off-target effects in a genome editing platform remain to be studied, sophisticated programmable nuclease-containing complexes and specialized nanoparticles have effectively improved the editing efficacies with minimal toxicity in mouse models. Specifically, lipid nanoparticles (LNPs), which are an FDA-approved approach for delivering siRNA to hepatocytes, have been developed and effectively exploited for sgRNAs and Cas9 mRNA delivery to mice, thereby allowing gene editing in not only the liver but also other organs [180,181]. Interestingly, Cas9 RNA- and sgRNA-packaged LNPs have been successfully used to treat patients suffering from transthyretin amyloidosis in an ongoing phase 1 clinical research [182]. A recent study demonstrated effective delivery of Cas9- or BE-RNPs by engineered DNA-free virus-like particles (eVLPs); this delivery method exhibits undetectable off-target effects and superiority compared to plasmid-, lentivirus-, or AAV-based delivery [183]. RNP-packaged eVLPs have been suggested as a favorable alternative to existing delivery approaches, owing to their transient exposure to gene editors, which limit off-target events [183]. A sponge-like silica nanoconstruct (SN) harboring Cas9- and BE-RNPs was recently developed for gene editing with low toxicity in mice [184]. Furthermore, improved genome editing via oviductal nucleic acid delivery (i-GONAD) has been widely utilized as an innovative approach to deliver genome editors directly into the murine fallopian tubes via electroporation [185,186,187]. However, insertion of a long DNA fragment using the i-GONAD scheme has been indicated to be less efficient compared to microinjection method [185]. This limitation was overcome with the Easi-CRISPR approach that enables efficient insertion of large DNA donors into genomes via microinjection of the RNPs and donors into mouse zygotes [188,189]. Interestingly, Gu et al. reported that co-injection of monomeric streptavidin-fused Cas9 mRNA, sgRNA, and biotinylated PCR donor DNA into two-cell stage embryos of mice (known as 2C-HR-CRISPR method) could effectively allow large insertions (up to 95%) at the target loci to minimize the undesirable effects on the integration sites [190,191]. In addition, CRISPR-READI is a recently developed method that combines delivery of CRISPR/Cas9 as an RNP via electroporation and donor DNA via an AAV system [192]; this method has been shown to allow successful insertion of large DNA fragments (up to 3.3 kb) into embryos [192]. Collectively, these recent breakthroughs and future advances in gene editing and delivery enable a broad spectrum of human disease models and therapeutic applications, which are expected to facilitate clinical transition and applications of promising programmable nucleases.

Author Contributions

H.T.L.P. and K.K. wrote the draft manuscript. H.T.L.P., K.K., H.L. and J.K.S. discussed and commented on the manuscript. K.K., H.L. and J.K.S. supervised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Bio & Medical Technology Development Program of the National Research Foundation (NRF) funded by the Korean government (NRF-2020M3A9D5A01082439 and 2019R1A2C2087198 to K.K., 2014M3A9D5A01075128 to H.L. and 2013M3A9D5072550 and 2021M3H9A1030158 to J.K.S.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data availability is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Simonetti, G.; Bertilaccio, M.T.S.; Ghia, P.; Klein, U. Mouse models in the study of chronic lymphocytic leukemia pathogenesis and therapy. Blood 2014, 124, 1010–1019. [Google Scholar] [CrossRef] [PubMed]
  2. Rosenthal, N.; Brown, S. The mouse ascending: Perspectives for human-disease models. Nat. Cell Biol. 2007, 9, 993–999. [Google Scholar] [CrossRef] [PubMed]
  3. Jaenisch, R.; Mintz, B. Simian Virus 40 DNA Sequences in DNA of Healthy Adult Mice Derived from Preimplantation Blastocysts Injected with Viral DNA. Proc. Natl. Acad. Sci. USA 1974, 71, 1250–1254. [Google Scholar] [CrossRef] [PubMed]
  4. Gaj, T.; Gersbach, C.A.; Barbas, C.F., III. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 2014, 31, 397–405. [Google Scholar] [CrossRef]
  5. Komor, A.C.; Kim, Y.B.; Packer, M.S.; Zuris, J.A.; Liu, D.R. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature 2016, 533, 420–424. [Google Scholar] [CrossRef] [PubMed]
  6. Anzalone, A.V.; Randolph, P.B.; Davis, J.R.; Sousa, A.A.; Koblan, L.W.; Levy, J.M.; Chen, P.J.; Wilson, C.; Newby, G.A.; Raguram, A.; et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature 2019, 576, 149–157. [Google Scholar] [CrossRef] [PubMed]
  7. Urnov, F.D.; Rebar, E.J.; Holmes, M.C.; Zhang, H.S.; Gregory, P.D. Genome editing with engineered zinc finger nucleases. Nat. Rev. Genet. 2010, 11, 636–646. [Google Scholar] [CrossRef]
  8. Becker, S.; Boch, J. TALE and TALEN genome editing technologies. Gene Genome Ed. 2021, 2, 100007. [Google Scholar] [CrossRef]
  9. Miller, J.C.; Holmes, M.C.; Wang, J.; Guschin, D.Y.; Lee, Y.-L.; Rupniewski, I.; Beausejour, C.M.; Waite, A.; Wang, N.S.; Kim, K.A.; et al. An improved zinc-finger nuclease architecture for highly specific genome editing. Nat. Biotechnol. 2007, 25, 778–785. [Google Scholar] [CrossRef]
  10. Carroll, D. Progress and prospects: Zinc-finger nucleases as gene therapy agents. Gene Ther. 2008, 15, 1463–1468. [Google Scholar] [CrossRef] [Green Version]
  11. Joung, J.K.; Sander, J.D. TALENs: A widely applicable technology for targeted genome editing. Nat. Rev. Mol. Cell Biol. 2013, 14, 49–55. [Google Scholar] [CrossRef]
  12. Yan, W.; Smith, C.; Cheng, L. Expanded activity of dimer nucleases by combining ZFN and TALEN for genome editing. Sci. Rep. 2013, 3, srep02376. [Google Scholar] [CrossRef]
  13. Mali, P.; Yang, L.; Esvelt, K.M.; Aach, J.; Guell, M.; DiCarlo, J.E.; Norville, J.E.; Church, G.M. RNA-Guided Human Genome Engineering via Cas9. Science 2013, 339, 823–826. [Google Scholar] [CrossRef]
  14. Cong, L.; Ran, F.A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P.D.; Wu, X.; Jiang, W.; Marraffini, L.A.; et al. Multiplex Genome Engineering Using CRISPR/Cas Systems. Science 2013, 339, 819–823. [Google Scholar] [CrossRef]
  15. Ayabe, S.; Nakashima, K.; Yoshiki, A. Off- and on-target effects of genome editing in mouse embryos. J. Reprod. Dev. 2019, 65, 1–5. [Google Scholar] [CrossRef]
  16. Zhang, X.-H.; Tee, L.Y.; Wang, X.-G.; Huang, Q.-S.; Yang, S.-H. Off-target effects in CRISPR/Cas9-mediated genome engineering. Mol. Ther. Nucleic Acids 2015, 4, e264. [Google Scholar] [CrossRef]
  17. Zetsche, B.; Gootenberg, J.S.; Abudayyeh, O.O.; Slaymaker, I.M.; Makarova, K.S.; Essletzbichler, P.; Volz, S.E.; Joung, J.; van der Oost, J.; Regev, A.; et al. Cpf1 Is a Single RNA-Guided Endonuclease of a Class 2 CRISPR-Cas System. Cell 2015, 163, 759–771. [Google Scholar] [CrossRef]
  18. Monteys, A.M.; Ebanks, S.A.; Keiser, M.S.; Davidson, B.L. CRISPR/Cas9 Editing of the Mutant Huntingtin Allele In Vitro and In Vivo. Mol. Ther. 2016, 25, 12–23. [Google Scholar] [CrossRef]
  19. Zhang, Y.; Long, C.; Li, H.; McAnally, J.R.; Baskin, K.K.; Shelton, J.M.; Bassel-Duby, R.; Olson, E.N. CRISPR-Cpf1 correction of muscular dystrophy mutations in human cardiomyocytes and mice. Sci. Adv. 2017, 3, e1602814. [Google Scholar] [CrossRef]
  20. Zhang, J.; Yang, F.; Yang, Y.; Jiang, Y.; Huo, Y.-X. Optimizing a CRISPR-Cpf1-based genome engineering system for Corynebacterium glutamicum. Microb. Cell Factories 2019, 18, 60. [Google Scholar] [CrossRef] [Green Version]
  21. Kim, D.; Kim, J.; Hur, J.K.; Been, K.W.; Yoon, S.-h.; Kim, J.-S. Genome-wide analysis reveals specificities of Cpf1 endonucleases in human cells. Nat. Biotechnol. 2016, 34, 863–868. [Google Scholar] [CrossRef] [PubMed]
  22. Abudayyeh, O.O.; Gootenberg, J.S.; Essletzbichler, P.; Han, S.; Joung, J.; Belanto, J.J.; Verdine, V.; Cox, D.B.T.; Kellner, M.J.; Regev, A.; et al. RNA targeting with CRISPR–Cas13. Nature 2017, 550, 280–284. [Google Scholar] [CrossRef] [PubMed]
  23. Konermann, S.; Lotfy, P.; Brideau, N.J.; Oki, J.; Shokhirev, M.N.; Hsu, P.D. Transcriptome Engineering with RNA-Targeting Type VI-D CRISPR Effectors. Cell 2018, 173, 665–676.e14. [Google Scholar] [CrossRef] [PubMed]
  24. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef] [PubMed]
  25. Cox, D.B.T.; Gootenberg, J.S.; Abudayyeh, O.O.; Franklin, B.; Kellner, M.J.; Joung, J.; Zhang, F. RNA editing with CRISPR-Cas13. Science 2017, 358, 1019–1027. [Google Scholar] [CrossRef]
  26. Yan, W.X.; Chong, S.; Zhang, H.; Makarova, K.S.; Koonin, E.V.; Cheng, D.R.; Scott, D.A. Cas13d Is a Compact RNA-Targeting Type VI CRISPR Effector Positively Modulated by a WYL-Domain-Containing Accessory Protein. Mol. Cell 2018, 70, 327–339.e5. [Google Scholar] [CrossRef]
  27. Du, M.; Jillette, N.; Zhu, J.J.; Li, S.; Cheng, A.W. CRISPR artificial splicing factors. Nat. Commun. 2020, 11, 1–11. [Google Scholar] [CrossRef]
  28. Mahas, A.; Aman, R.; Mahfouz, M. CRISPR-Cas13d mediates robust RNA virus interference in plants. Genome Biol. 2019, 20, 1–16. [Google Scholar] [CrossRef]
  29. Yu, Y.; Pan, Z.; Wang, X.; Bian, X.; Wang, W.; Liang, Q.; Kou, M.; Ji, H.; Li, Y.; Ma, D.; et al. Targeting of SPCSV- RNase3 via CRISPR-Cas13 confers resistance against sweet potato virus disease. Mol. Plant Pathol. 2021, 23, 104–117. [Google Scholar] [CrossRef]
  30. Newby, G.A.; Liu, D.R. In vivo somatic cell base editing and prime editing. Mol. Ther. 2021, 29, 3107–3124. [Google Scholar] [CrossRef]
  31. Chen, P.J.; Liu, D.R. Prime editing for precise and highly versatile genome manipulation. Nat. Rev. Genet. 2022, 1–17. [Google Scholar] [CrossRef]
  32. Huang, T.P.; Newby, G.A.; Liu, D.R. Precision genome editing using cytosine and adenine base editors in mammalian cells. Nat. Protoc. 2021, 16, 1089–1128. [Google Scholar] [CrossRef]
  33. Anzalone, A.V.; Koblan, L.; Liu, D.R. Genome editing with CRISPR–Cas nucleases, base editors, transposases and prime editors. Nat. Biotechnol. 2020, 38, 824–844. [Google Scholar] [CrossRef]
  34. Liu, Y.; Li, X.; He, S.; Huang, S.; Li, C.; Chen, Y.; Liu, Z.; Huang, X.; Wang, X. Efficient generation of mouse models with the prime editing system. Cell Discov. 2020, 6, 1–4. [Google Scholar] [CrossRef]
  35. Caso, F.; Davies, B. Base editing and prime editing in laboratory animals. Lab. Anim. 2021, 56, 35–49. [Google Scholar] [CrossRef]
  36. Carbery, I.D.; Ji, D.; Harrington, A.; Brown, V.; Weinstein, E.J.; Liaw, L.; Cui, X. Targeted Genome Modification in Mice Using Zinc-Finger Nucleases. Genetics 2010, 186, 451–459. [Google Scholar] [CrossRef]
  37. Hermann, M.; Maeder, M.L.; Rector, K.; Ruiz, J.; Becher, B.; Bürki, K.; Khayter, C.; Aguzzi, A.; Joung, J.K.; Buch, T.; et al. Evaluation of OPEN Zinc Finger Nucleases for Direct Gene Targeting of the ROSA26 Locus in Mouse Embryos. PLoS ONE 2012, 7, e41796. [Google Scholar] [CrossRef]
  38. Sung, Y.H.; Baek, I.-J.; Kim, D.H.; Jeon, J.; Lee, J.; Lee, K.; Jeong, D.; Kim, J.-S.; Lee, H.-W. Knockout mice created by TALEN-mediated gene targeting. Nat. Biotechnol. 2013, 31, 23–24. [Google Scholar] [CrossRef]
  39. Takada, S.; Sato, T.; Ito, Y.; Yamashita, S.; Kato, T.; Kawasumi, M.; Kanai-Azuma, M.; Igarashi, A.; Kato, T.; Tamano, M.; et al. Targeted Gene Deletion of miRNAs in Mice by TALEN System. PLoS ONE 2013, 8, e76004. [Google Scholar] [CrossRef]
  40. Ma, K.; Wang, J.; Shen, B.; Qiu, L.; Huang, X.; Li, Z. Efficient targeting of FATS at a common fragile site in mice through TALEN-mediated double-hit genome modification. Biotechnol. Lett. 2013, 36, 471–479. [Google Scholar] [CrossRef]
  41. Liu, Y.; Lv, X.; Tan, R.; Liu, T.; Chen, T.; Li, M.; Liu, Y.; Nie, F.; Wang, X.; Zhou, P.; et al. A Modified TALEN-Based Strategy for Rapidly and Efficiently Generating Knockout Mice for Kidney Development Studies. PLoS ONE 2014, 9, e84893. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, H.; Yang, H.; Shivalila, C.S.; Dawlaty, M.M.; Cheng, A.W.; Zhang, F.; Jaenisch, R. One-Step Generation of Mice Carrying Mutations in Multiple Genes by CRISPR/Cas-Mediated Genome Engineering. Cell 2013, 153, 910–918. [Google Scholar] [CrossRef] [PubMed]
  43. Meyer, M.; de Angelis, M.H.; Wurst, W.; Kühn, R. Gene targeting by homologous recombination in mouse zygotes mediated by zinc-finger nucleases. Proc. Natl. Acad. Sci. USA 2010, 107, 15022–15026. [Google Scholar] [CrossRef] [PubMed]
  44. An, L.; Hu, Y.; Chang, S.; Zhu, X.; Ling, P.; Zhang, F.; Liu, J.; Liu, Y.; Chen, Y.; Yang, L.; et al. Efficient generation of FVII gene knockout mice using CRISPR/Cas9 nuclease and truncated guided RNAs. Sci. Rep. 2016, 6, 25199. [Google Scholar] [CrossRef] [PubMed]
  45. Davies, B.; Davies, G.; Preece, C.; Puliyadi, R.; Szumska, D.; Bhattacharya, S. Site Specific Mutation of the Zic2 Locus by Microinjection of TALEN mRNA in Mouse CD1, C3H and C57BL/6J Oocytes. PLoS ONE 2013, 8, e60216. [Google Scholar] [CrossRef]
  46. Qiu, Z.; Liu, M.; Chen, Z.; Shao, Y.; Pan, H.; Wei, G.; Yu, C.; Zhang, L.; Li, X.; Wang, P.; et al. High-efficiency and heritable gene targeting in mouse by transcription activator-like effector nucleases. Nucleic Acids Res. 2013, 41, e120. [Google Scholar] [CrossRef]
  47. Wefers, B.; Panda, S.K.; Ortiz, O.; Brandl, C.; Hensler, S.; Hansen, J.; Wurst, W.; Kühn, R. Generation of targeted mouse mutants by embryo microinjection of TALEN mRNA. Nat. Protoc. 2013, 8, 2355–2379. [Google Scholar] [CrossRef]
  48. Zhou, J.; Shen, B.; Zhang, W.; Wang, J.; Yang, J.; Chen, L.; Zhang, N.; Zhu, K.; Xu, J.; Hu, B.; et al. One-step generation of different immunodeficient mice with multiple gene modifications by CRISPR/Cas9 mediated genome engineering. Int. J. Biochem. Cell Biol. 2014, 46, 49–55. [Google Scholar] [CrossRef]
  49. Mashiko, D.; Young, S.A.M.; Muto, M.; Kato, H.; Nozawa, K.; Ogawa, M.; Noda, T.; Kim, Y.-J.; Satouh, Y.; Fujihara, Y.; et al. Feasibility for a large scale mouse mutagenesis by injecting CRISPR/Cas plasmid into zygotes. Dev. Growth Differ. 2013, 56, 122–129. [Google Scholar] [CrossRef]
  50. Lemoine, A.; Friec, G.C.-L.; Langa, F.; Louvet, C. Generation of a Double KO Mouse by Simultaneous Targeting of the Neighboring Genes Tmem176a and Tmem176b Using CRISPR/Cas9: Key Steps from Design to Genotyping. J. Genet. Genom. 2016, 43, 329–340. [Google Scholar] [CrossRef]
  51. Boroviak, K.; Doe, B.; Banerjee, R.; Yang, F.; Bradley, A. Chromosome engineering in zygotes with CRISPR/C as9. Genesis 2016, 54, 78–85. [Google Scholar] [CrossRef]
  52. Sung, Y.H.; Jin, Y.; Kim, S.; Lee, H.-W. Generation of knockout mice using engineered nucleases. Methods 2014, 69, 85–93. [Google Scholar] [CrossRef]
  53. Shen, B.; Zhang, J.; Wu, H.; Wang, J.; Ma, K.; Li, Z.; Zhang, X.; Zhang, P.; Huang, X. Generation of gene-modified mice via Cas9/RNA-mediated gene targeting. Cell Res. 2013, 23, 720–723. [Google Scholar] [CrossRef]
  54. Horii, T.; Arai, Y.; Yamazaki, M.; Morita, S.; Kimura, M.; Itoh, M.; Abe, Y.; Hatada, I. Validation of microinjection methods for generating knockout mice by CRISPR/Cas-mediated genome engineering. Sci. Rep. 2014, 4, 4513. [Google Scholar] [CrossRef]
  55. Li, F.; Cowley, D.O.; Banner, D.; Holle, E.; Zhang, L.; Su, L. Efficient genetic manipulation of the NOD-Rag1-/-IL2RgammaC-null mouse by combining in vitro fertilization and CRISPR/Cas9 technology. Sci. Rep. 2014, 4, 5290. [Google Scholar] [CrossRef]
  56. Duan, X.; Liu, J.; Zheng, X.; Wang, Z.; Zhang, Y.; Hao, Y.; Yang, T.; Deng, H. Deficiency of ATP6V1H Causes Bone Loss by Inhibiting Bone Resorption and Bone Formation through the TGF-β1 Pathway. Theranostics 2016, 6, 2183–2195. [Google Scholar] [CrossRef]
  57. Canalis, E.; Yu, J.; Schilling, L.; Yee, S.-P.; Zanotti, S. The lateral meningocele syndrome mutation causes marked osteopenia in mice. J. Biol. Chem. 2018, 293, 14165–14177. [Google Scholar] [CrossRef]
  58. Pankowicz, F.P.; Barzi, M.; Legras, X.; Hubert, L.; Mi, T.; Tomolonis, J.A.; Ravishankar, M.; Sun, Q.; Yang, D.; Borowiak, M.; et al. Reprogramming metabolic pathways in vivo with CRISPR/Cas9 genome editing to treat hereditary tyrosinaemia. Nat. Commun. 2016, 7, 12642. [Google Scholar] [CrossRef]
  59. Pankowicz, F.P.; Barzi, M.; Kim, K.H.; Legras, X.; Martins, C.S.; Wooton-Kee, C.R.; Lagor, W.R.; Marini, J.C.; Elsea, S.; Bissig-Choisat, B.; et al. Rapid Disruption of Genes Specifically in Livers of Mice Using Multiplex CRISPR/Cas9 Editing. Gastroenterology 2018, 155, 1967–1970.e6. [Google Scholar] [CrossRef]
  60. Kato, T.; Hara, S.; Goto, Y.; Ogawa, Y.; Okayasu, H.; Kubota, S.; Tamano, M.; Terao, M.; Takada, S. Creation of mutant mice with megabase-sized deletions containing custom-designed breakpoints by means of the CRISPR/Cas9 system. Sci. Rep. 2017, 7, 9. [Google Scholar] [CrossRef] [Green Version]
  61. Hara, S.; Kato, T.; Goto, Y.; Kubota, S.; Tamano, M.; Terao, M.; Takada, S. Microinjection-based generation of mutant mice with a double mutation and a 0.5 Mb deletion in their genome by the CRISPR/Cas9 system. J. Reprod. Dev. 2016, 62, 531–536. [Google Scholar] [CrossRef] [PubMed]
  62. Li, J.; Shou, J.; Guo, Y.; Tang, Y.; Wu, Y.; Jia, Z.; Zhai, Y.; Chen, Z.; Xu, Q.; Wu, Q. Efficient inversions and duplications of mammalian regulatory DNA elements and gene clusters by CRISPR/Cas9. J. Mol. Cell Biol. 2015, 7, 284–298. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, L.; Shao, Y.; Guan, Y.; Li, L.; Wu, L.; Chen, F.; Liu, M.; Chen, H.; Ma, Y.; Ma, X.; et al. Large genomic fragment deletion and functional gene cassette knock-in via Cas9 protein mediated genome editing in one-cell rodent embryos. Sci. Rep. 2015, 5, 17517. [Google Scholar] [CrossRef] [PubMed]
  64. Mizuno, S.; Takami, K.; Daitoku, Y.; Tanimoto, Y.; Dinh, T.T.H.; Mizuno-Iijima, S.; Hasegawa, Y.; Takahashi, S.; Sugiyama, F.; Yagami, K.-I. Peri-implantation lethality in mice carrying megabase-scale deletion on 5qc3.3 is caused by Exoc1 null mutation. Sci. Rep. 2015, 5, 13632. [Google Scholar] [CrossRef]
  65. Mochizuki, Y.; Chiba, T.; Kataoka, K.; Yamashita, S.; Sato, T.; Kato, T.; Takahashi, K.; Miyamoto, T.; Kitazawa, M.; Hatta, T.; et al. Combinatorial CRISPR/Cas9 Approach to Elucidate a Far-Upstream Enhancer Complex for Tissue-Specific Sox9 Expression. Dev. Cell 2018, 46, 794–806.e6. [Google Scholar] [CrossRef]
  66. Platt, R.J.; Chen, S.; Zhou, Y.; Yim, M.J.; Swiech, L.; Kempton, H.R.; Dahlman, J.E.; Parnas, O.; Eisenhaure, T.M.; Jovanovic, M.; et al. CRISPR-Cas9 Knockin Mice for Genome Editing and Cancer Modeling. Cell 2014, 159, 440–455. [Google Scholar] [CrossRef]
  67. Kim, K.; Ryu, S.-M.; Kim, S.-T.; Baek, G.; Kim, D.; Lim, K.; Chung, E.; Kim, S.; Kim, J.-S. Highly efficient RNA-guided base editing in mouse embryos. Nat. Biotechnol. 2017, 35, 435–437. [Google Scholar] [CrossRef]
  68. Zhang, H.; Pan, H.; Zhou, C.; Wei, Y.; Ying, W.; Li, S.; Wang, G.; Li, C.; Ren, Y.; Li, G.; et al. Simultaneous zygotic inactivation of multiple genes in mouse through CRISPR/Cas9-mediated base editing. Development 2018, 145, dev.168906. [Google Scholar] [CrossRef]
  69. Wang, X.; Liu, Z.; Li, G.; Dang, L.; Huang, S.; He, L.; Ma, Y.; Li, C.; Liu, M.; Yang, G.; et al. Efficient Gene Silencing by Adenine Base Editor-Mediated Start Codon Mutation. Mol. Ther. 2019, 28, 431–440. [Google Scholar] [CrossRef]
  70. Lin, J.; Liu, X.; Lu, Z.; Huang, S.; Wu, S.; Yu, W.; Liu, Y.; Zheng, X.; Huang, X.; Sun, Q.; et al. Modeling a cataract disorder in mice with prime editing. Mol. Ther. - Nucleic Acids 2021, 25, 494–501. [Google Scholar] [CrossRef]
  71. Meyer, M.; Ortiz, O.; de Angelis, M.H.; Wurst, W.; Kühn, R. Modeling disease mutations by gene targeting in one-cell mouse embryos. Proc. Natl. Acad. Sci. USA 2012, 109, 9354–9359. [Google Scholar] [CrossRef]
  72. Wefers, B.; Meyer, M.; Ortiz, O.; de Angelis, M.H.; Hansen, J.; Wurst, W.; Kühn, R. Direct production of mouse disease models by embryo microinjection of TALENs and oligodeoxynucleotides. Proc. Natl. Acad. Sci. USA 2013, 110, 3782–3787. [Google Scholar] [CrossRef]
  73. Panda, S.K.; Wefers, B.; Ortiz, O.; Floss, T.; Schmid, B.; Haass, C.; Wurst, W.; Kühn, R. Highly Efficient Targeted Mutagenesis in Mice Using TALENs. Genetics 2013, 195, 703–713. [Google Scholar] [CrossRef]
  74. Low, B.E.; Krebs, M.P.; Joung, J.K.; Tsai, S.Q.; Nishina, P.M.; Wiles, M.V. Correction of the Crb1rd8 Allele and Retinal Phenotype in C57BL/6N Mice Via TALEN-Mediated Homology-Directed Repair. Investig. Opthalmol. Vis. Sci. 2014, 55, 387–395. [Google Scholar] [CrossRef]
  75. Jones, J.M.; Meisler, M.H. Modeling human epilepsy by TALEN targeting of mouse sodium channel Scn8a. Genesis 2013, 52, 141–148. [Google Scholar] [CrossRef]
  76. Yang, H.; Wang, H.; Shivalila, C.S.; Cheng, A.W.; Shi, L.; Jaenisch, R. One-Step Generation of Mice Carrying Reporter and Conditional Alleles by CRISPR/Cas-Mediated Genome Engineering. Cell 2013, 154, 1370–1379. [Google Scholar] [CrossRef]
  77. Aida, T.; Chiyo, K.; Usami, T.; Ishikubo, H.; Imahashi, R.; Wada, Y.; Tanaka, K.F.; Sakuma, T.; Yamamoto, T.; Tanaka, K. Cloning-free CRISPR/Cas system facilitates functional cassette knock-in in mice. Genome Biol. 2015, 16, 1–11. [Google Scholar] [CrossRef]
  78. Rauch, F.; Geng, Y.; Lamplugh, L.; Hekmatnejad, B.; Gaumond, M.-H.; Penney, J.; Yamanaka, Y.; Moffatt, P. Crispr-Cas9 engineered osteogenesis imperfecta type V leads to severe skeletal deformities and perinatal lethality in mice. Bone 2018, 107, 131–142. [Google Scholar] [CrossRef]
  79. Yang, S.; Yang, H.; Huang, L.; Chen, L.; Qin, Z.; Li, S.; Li, X.-J. Lack of RAN-mediated toxicity in Huntington’s disease knock-in mice. Proc. Natl. Acad. Sci. USA 2020, 117, 4411–4417. [Google Scholar] [CrossRef]
  80. Hofer, U.; Henley, J.E.; Exline, C.M.; Mulhern, O.; Lopez, E.; Cannon, P.M. Pre-clinical Modeling of CCR5 Knockout in Human Hematopoietic Stem Cells by Zinc Finger Nucleases Using Humanized Mice. J. Infect. Dis. 2013, 208, S160–S164. [Google Scholar] [CrossRef] [Green Version]
  81. Holt, N.; Wang, J.; Kim, K.; Friedman, G.; Wang, X.; Taupin, V.; Crooks, G.M.; Kohn, D.B.; Gregory, P.D.; Holmes, M.C.; et al. Human hematopoietic stem/progenitor cells modified by zinc-finger nucleases targeted to CCR5 control HIV-1 in vivo. Nat. Biotechnol. 2010, 28, 839–847. [Google Scholar] [CrossRef] [PubMed]
  82. Li, L.; Krymskaya, L.; Wang, J.; Henley, J.; Rao, A.; Cao, L.-F.; Tran, C.-A.; Torres-Coronado, M.; Gardner, A.; Gonzalez, N.; et al. Genomic Editing of the HIV-1 Coreceptor CCR5 in Adult Hematopoietic Stem and Progenitor Cells Using Zinc Finger Nucleases. Mol. Ther. 2013, 21, 1259–1269. [Google Scholar] [CrossRef] [PubMed]
  83. Lux, C.; Pattabhi, S.; Berger, M.; Nourigat, C.; Flowers, D.A.; Negre, O.; Humbert, O.; Yang, J.G.; Lee, C.; Jacoby, K.; et al. TALEN-Mediated Gene Editing of HBG in Human Hematopoietic Stem Cells Leads to Therapeutic Fetal Hemoglobin Induction. Mol. Ther. Methods Clin. Dev. 2018, 12, 175–183. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, Y.; Li, H.; Nishiyama, T.; McAnally, J.R.; Sanchez-Ortiz, E.; Huang, J.; Mammen, P.P.; Bassel-Duby, R.; Olson, E.N. A humanized knockin mouse model of Duchenne muscular dystrophy and its correction by CRISPR-Cas9 therapeutic gene editing. Mol. Ther. Nucleic Acids 2022, 29, 525–537. [Google Scholar] [CrossRef]
  85. Burova, E.; Hermann, A.; Dai, J.; Ullman, E.; Halasz, G.; Potocky, T.; Hong, S.; Liu, M.; Allbritton, O.; Woodruff, A.; et al. Preclinical Development of the Anti-LAG-3 Antibody REGN3767: Characterization and Activity in Combination with the Anti-PD-1 Antibody Cemiplimab in Human PD-1xLAG-3–Knockin Mice. Mol. Cancer Ther. 2019, 18, 2051–2062. [Google Scholar] [CrossRef]
  86. Cui, X.; Ji, D.; A Fisher, D.; Wu, Y.; Briner, D.M.; Weinstein, E.J. Targeted integration in rat and mouse embryos with zinc-finger nucleases. Nat. Biotechnol. 2011, 29, 64–67. [Google Scholar] [CrossRef]
  87. Sommer, D.; Peters, A.E.; Wirtz, T.; Mai, M.; Ackermann, J.; Thabet, Y.; Schmidt, J.; Weighardt, H.; Wunderlich, F.T.; Degen, J.; et al. Efficient genome engineering by targeted homologous recombination in mouse embryos using transcription activator-like effector nucleases. Nat. Commun. 2014, 5, 3045. [Google Scholar] [CrossRef]
  88. Tabrizi, S.J.; Flower, M.; Ross, C.A.; Wild, E.J. Huntington disease: New insights into molecular pathogenesis and therapeutic opportunities. Nat. Rev. Neurol. 2020, 16, 1–18. [Google Scholar] [CrossRef]
  89. Serfecz, J.; Bazick, H.; Al Salihi, M.O.; Turner, P.; Fields, C.; Cruz, P.; Renne, R.; Notterpek, L. Downregulation of the human peripheral myelin protein 22 gene by miR-29a in cellular models of Charcot–Marie–Tooth disease. Gene Ther. 2019, 26, 455–464. [Google Scholar] [CrossRef]
  90. Caron, N.S.; Southwell, A.L.; Brouwers, C.C.; Cengio, L.D.; Xie, Y.; Black, H.F.; Anderson, L.M.; Ko, S.; Zhu, X.; Van Deventer, S.J.; et al. Potent and sustained huntingtin lowering via AAV5 encoding miRNA preserves striatal volume and cognitive function in a humanized mouse model of Huntington disease. Nucleic Acids Res. 2019, 48, 36–54. [Google Scholar] [CrossRef]
  91. Allen, T.; Brehm, M.; Bridges, S.; Ferguson, S.; Kumar, P.; Mirochnitchenko, O.; Palucka, K.; Pelanda, R.; Sanders-Beer, B.; Shultz, L.D.; et al. Humanized immune system mouse models: Progress, challenges and opportunities. Nat. Immunol. 2019, 20, 770–774. [Google Scholar] [CrossRef]
  92. Verhaart, I.E.C.; Aartsma-Rus, A. Therapeutic developments for Duchenne muscular dystrophy. Nat. Rev. Neurol. 2019, 15, 373–386. [Google Scholar] [CrossRef]
  93. Cho, A.; Haruyama, N.; Kulkarni, A.B. Generation of Transgenic Mice. Curr. Protoc. Cell Biol. 2009, 42, 19.11.1–19.11.22. [Google Scholar] [CrossRef]
  94. Chiang, C.; Jacobsen, J.C.; Ernst, C.; Hanscom, C.; Heilbut, A.; Blumenthal, I.; Mills, R.; Kirby, A.; Lindgren, A.M.; Rudiger, S.R.; et al. Complex reorganization and predominant non-homologous repair following chromosomal breakage in karyotypically balanced germline rearrangements and transgenic integration. Nat. Genet. 2012, 44, 390–397. [Google Scholar] [CrossRef]
  95. Gemberling, M.P.; Siklenka, K.; Rodriguez, E.; Tonn-Eisinger, K.R.; Barrera, A.; Liu, F.; Kantor, A.; Li, L.; Cigliola, V.; Hazlett, M.F.; et al. Transgenic mice for in vivo epigenome editing with CRISPR-based systems. Nat. Methods 2021, 18, 965–974. [Google Scholar] [CrossRef]
  96. Xu, L.; Park, K.H.; Zhao, L.; Xu, J.; El Refaey, M.; Gao, Y.; Zhu, H.; Ma, J.; Han, R. CRISPR-mediated Genome Editing Restores Dystrophin Expression and Function in mdx Mice. Mol. Ther. 2016, 24, 564–569. [Google Scholar] [CrossRef]
  97. Bengtsson, N.E.; Hall, J.K.; Odom, G.L.; Phelps, M.P.; Andrus, C.R.; Hawkins, R.D.; Hauschka, S.D.; Chamberlain, J.R.; Chamberlain, J.S. Muscle-specific CRISPR/Cas9 dystrophin gene editing ameliorates pathophysiology in a mouse model for Duchenne muscular dystrophy. Nat. Commun. 2017, 8, 14454. [Google Scholar] [CrossRef]
  98. Nelson, C.E.; Hakim, C.H.; Ousterout, D.G.; Thakore, P.I.; Moreb, E.A.; Rivera, R.M.C.; Madhavan, S.; Pan, X.; Ran, F.A.; Yan, W.X.; et al. In vivo genome editing improves muscle function in a mouse model of Duchenne muscular dystrophy. Science 2016, 351, 403–407. [Google Scholar] [CrossRef]
  99. Long, C.; McAnally, J.R.; Shelton, J.M.; Mireault, A.A.; Bassel-Duby, R.; Olson, E.N. Prevention of muscular dystrophy in mice by CRISPR/Cas9–mediated editing of germline DNA. Science 2014, 345, 1184–1188. [Google Scholar] [CrossRef]
  100. Tabebordbar, M.; Zhu, K.; Cheng, J.K.W.; Chew, W.L.; Widrick, J.J.; Yan, W.X.; Maesner, C.; Wu, E.Y.; Xiao, R.; Ran, F.A.; et al. In vivo gene editing in dystrophic mouse muscle and muscle stem cells. Science 2016, 351, 407–411. [Google Scholar] [CrossRef] [Green Version]
  101. Amoasii, L.; Long, C.; Li, H.; Mireault, A.A.; Shelton, J.M.; Sanchez-Ortiz, E.; McAnally, J.R.; Bhattacharyya, S.; Schmidt, F.; Grimm, D.; et al. Single-cut genome editing restores dystrophin expression in a new mouse model of muscular dystrophy. Sci. Transl. Med. 2017, 9, e8081. [Google Scholar] [CrossRef] [PubMed]
  102. Min, Y.-L.; Li, H.; Rodriguez-Caycedo, C.; Mireault, A.A.; Huang, J.; Shelton, J.M.; McAnally, J.R.; Amoasii, L.; Mammen, P.P.A.; Bassel-Duby, R.; et al. CRISPR-Cas9 corrects Duchenne muscular dystrophy exon 44 deletion mutations in mice and human cells. Sci. Adv. 2019, 5, eaav4324. [Google Scholar] [CrossRef] [PubMed]
  103. Min, Y.-L.; Chemello, F.; Li, H.; Rodriguez-Caycedo, C.; Sanchez-Ortiz, E.; Mireault, A.A.; McAnally, J.R.; Shelton, J.M.; Zhang, Y.; Bassel-Duby, R.; et al. Correction of Three Prominent Mutations in Mouse and Human Models of Duchenne Muscular Dystrophy by Single-Cut Genome Editing. Mol. Ther. 2020, 28, 2044–2055. [Google Scholar] [CrossRef] [PubMed]
  104. Chemello, F.; Chai, A.C.; Li, H.; Rodriguez-Caycedo, C.; Sanchez-Ortiz, E.; Atmanli, A.; Mireault, A.A.; Liu, N.; Bassel-Duby, R.; Olson, E.N. Precise correction of Duchenne muscular dystrophy exon deletion mutations by base and prime editing. Sci. Adv. 2021, 7, e4910. [Google Scholar] [CrossRef]
  105. Nelson, C.; Wu, Y.; Gemberling, M.P.; Oliver, M.L.; Waller, M.A.; Bohning, J.D.; Robinson-Hamm, J.N.; Bulaklak, K.; Rivera, R.M.C.; Collier, J.H.; et al. Long-term evaluation of AAV-CRISPR genome editing for Duchenne muscular dystrophy. Nat. Med. 2019, 25, 427–432. [Google Scholar] [CrossRef]
  106. Ryu, S.-M.; Koo, T.; Kim, K.; Lim, K.; Baek, G.; Kim, S.-T.; Kim, H.S.; Kim, D.-E.; Lee, H.; Chung, E.; et al. Adenine base editing in mouse embryos and an adult mouse model of Duchenne muscular dystrophy. Nat. Biotechnol. 2018, 36, 536–539. [Google Scholar] [CrossRef]
  107. Xu, L.; Zhang, C.; Li, H.; Wang, P.; Gao, Y.; Mokadam, N.A.; Ma, J.; Arnold, W.D.; Han, R. Efficient precise in vivo base editing in adult dystrophic mice. Nat. Commun. 2021, 12, e23996. [Google Scholar] [CrossRef]
  108. Wei, T.; Cheng, Q.; Min, Y.-L.; Olson, E.N.; Siegwart, D.J. Systemic nanoparticle delivery of CRISPR-Cas9 ribonucleoproteins for effective tissue specific genome editing. Nat. Commun. 2020, 11, 3232. [Google Scholar] [CrossRef]
  109. Domenig, S.A.; Bundschuh, N.; Lenardič, A.; Ghosh, A.; Kim, I.; Qabrati, X.; D’Hulst, G.; Bar-Nur, O. CRISPR/Cas9 editing of directly reprogrammed myogenic progenitors restores dystrophin expression in a mouse model of muscular dystrophy. Stem Cell Rep. 2022, 17, 321–336. [Google Scholar] [CrossRef]
  110. Chen, H.; Shi, M.; Gilam, A.; Zheng, Q.; Zhang, Y.; Afrikanova, I.; Li, J.; Gluzman, Z.; Jiang, R.; Kong, L.-J.; et al. Hemophilia A ameliorated in mice by CRISPR-based in vivo genome editing of human Factor VIII. Sci. Rep. 2019, 9, e53198. [Google Scholar] [CrossRef] [Green Version]
  111. Luo, S.; Li, Z.; Dai, X.; Zhang, R.; Liang, Z.; Li, W.; Zeng, M.; Su, J.; Wang, J.; Liang, X.; et al. CRISPR/Cas9-Mediated in vivo Genetic Correction in a Mouse Model of Hemophilia A. Front. Cell Dev. Biol. 2021, 9. [Google Scholar] [CrossRef]
  112. Zhang, J.-P.; Cheng, X.-X.; Zhao, M.; Li, G.-H.; Xu, J.; Zhang, F.; Yin, M.-D.; Meng, F.-Y.; Dai, X.-Y.; Fu, Y.-W.; et al. Curing hemophilia A by NHEJ-mediated ectopic F8 insertion in the mouse. Genome Biol. 2019, 20, e1907. [Google Scholar] [CrossRef]
  113. Li, H.; Haurigot, V.; Doyon, Y.; Li, T.; Wong, S.Y.; Bhagwat, A.S.; Malani, N.; Anguela, X.M.; Sharma, R.; Ivanciu, L.; et al. In vivo genome editing restores haemostasis in a mouse model of haemophilia. Nature 2011, 475, 217–221. [Google Scholar] [CrossRef]
  114. Anguela, X.M.; Sharma, R.; Doyon, Y.; Miller, J.C.; Li, H.; Haurigot, V.A.; Rohde, M.E.; Wong, S.Y.; Davidson, R.J.; Zhou, S.; et al. Robust ZFN-mediated genome editing in adult hemophilic mice. Blood 2013, 122, 3283–3287. [Google Scholar] [CrossRef]
  115. Guan, Y.; Ma, Y.; Li, Q.; Sun, Z.; Ma, L.; Wu, L.; Wang, L.; Zeng, L.; Shao, Y.; Chen, Y.; et al. CRISPR /Cas9-mediated somatic correction of a novel coagulator factor IX gene mutation ameliorates hemophilia in mouse. EMBO Mol. Med. 2016, 8, 477–488. [Google Scholar] [CrossRef]
  116. Lee, J.H.; Oh, H.-K.; Choi, B.S.; Lee, H.H.; Lee, K.J.; Kim, U.G.; Lee, J.; Lee, H.; Lee, G.S.; Ahn, S.J.; et al. Genome editing-mediated knock-in of therapeutic genes ameliorates the disease phenotype in a model of hemophilia. Mol. Ther. Nucleic Acids 2022, 29, 551–562. [Google Scholar] [CrossRef]
  117. Chen, X.; Niu, X.; Liu, Y.; Zheng, R.; Yang, L.; Lu, J.; Yin, S.; Wei, Y.; Pan, J.; Sayed, A.; et al. Long-term correction of hemophilia B through CRISPR/Cas9 induced homology-independent targeted integration. J. Genet. Genom. 2022, 49, 1114–1126. [Google Scholar] [CrossRef]
  118. Wang, Q.; Zhong, X.; Li, Q.; Su, J.; Liu, Y.; Mo, L.; Deng, H.; Yang, Y. CRISPR-Cas9-Mediated In Vivo Gene Integration at the Albumin Locus Recovers Hemostasis in Neonatal and Adult Hemophilia B Mice. Mol. Ther. Methods Clin. Dev. 2020, 18, 520–531. [Google Scholar] [CrossRef]
  119. Garriga-Canut, M.; Agustín-Pavón, C.; Herrmann, F.; Sánchez, A.; Dierssen, M.; Fillat, C.; Isalan, M. Synthetic zinc finger repressors reduce mutant huntingtin expression in the brain of R6/2 mice. Proc. Natl. Acad. Sci. USA 2012, 109, E3136–E3145. [Google Scholar] [CrossRef]
  120. Liu, H.; Zhang, C.; Xu, J.; Jin, J.; Cheng, L.; Miao, X.; Wu, Q.; Wei, Z.; Liu, P.; Lu, H.; et al. Huntingtin silencing delays onset and slows progression of Huntington’s disease: A biomarker study. Brain 2021, 144, 3101–3113. [Google Scholar] [CrossRef]
  121. Ekman, F.K.; Ojala, D.S.; Adil, M.M.; Lopez, P.A.; Schaffer, D.V.; Gaj, T. CRISPR-Cas9-Mediated Genome Editing Increases Lifespan and Improves Motor Deficits in a Huntington’s Disease Mouse Model. Mol. Ther. Nucleic Acids 2019, 17, 829–839. [Google Scholar] [CrossRef]
  122. Yang, S.; Chang, R.; Yang, H.; Zhao, T.; Hong, Y.; Kong, H.E.; Sun, X.; Qin, Z.; Jin, P.; Li, S.; et al. CRISPR/Cas9-mediated gene editing ameliorates neurotoxicity in mouse model of Huntington’s disease. J. Clin. Investig. 2017, 127, 2719–2724. [Google Scholar] [CrossRef] [PubMed]
  123. Lim, C.K.W.; McCallister, T.X.; Saporito-Magriña, C.; McPheron, G.D.; Krishnan, R.; Zeballos C, M.A.; Powell, J.E.; Clark, L.V.; Perez-Pinera, P.; Gaj, T. CRISPR base editing of cis-regulatory elements enables the perturbation of neurodegeneration-linked genes. Mol. Ther. 2022, 30, 3619–3631. [Google Scholar] [CrossRef] [PubMed]
  124. Powell, J.E.; Lim, C.K.W.; Krishnan, R.; McCallister, T.X.; Saporito-Magriña, C.; Zeballos, M.A.; McPheron, G.D.; Gaj, T. Targeted gene silencing in the nervous system with CRISPR-Cas13. Sci. Adv. 2022, 8, e2485. [Google Scholar] [CrossRef] [PubMed]
  125. Gaj, T.; Ojala, D.S.; Ekman, F.K.; Byrne, L.C.; Limsirichai, P.; Schaffer, D.V. In vivo genome editing improves motor function and extends survival in a mouse model of ALS. Sci. Adv. 2017, 3, eaar3952. [Google Scholar] [CrossRef]
  126. Duan, W.; Guo, M.; Yi, L.; Liu, Y.; Li, Z.; Ma, Y.; Zhang, G.; Liu, Y.; Bu, H.; Song, X.; et al. The deletion of mutant SOD1 via CRISPR/Cas9/sgRNA prolongs survival in an amyotrophic lateral sclerosis mouse model. Gene Ther. 2020, 27, 157–169. [Google Scholar] [CrossRef]
  127. Lim, C.K.; Gapinske, M.; Brooks, A.K.; Woods, W.S.; Powell, J.E.; C., M.A.Z.; Winter, J.; Perez-Pinera, P.; Gaj, T. Treatment of a Mouse Model of ALS by In Vivo Base Editing. Mol. Ther. 2020, 28, 1177–1189. [Google Scholar] [CrossRef]
  128. Gao, X.; Tao, Y.; Lamas, V.; Huang, M.; Yeh, W.-H.; Pan, B.; Hu, Y.-J.; Hu, J.H.; Thompson, D.B.; Shu, Y.; et al. Treatment of autosomal dominant hearing loss by in vivo delivery of genome editing agents. Nature 2018, 553, 217–221. [Google Scholar] [CrossRef]
  129. Yeh, W.-H.; Shubina-Oleinik, O.; Levy, J.M.; Pan, B.; Newby, G.A.; Wornow, M.; Burt, R.; Chen, J.C.; Holt, J.R.; Liu, D.R. In vivo base editing restores sensory transduction and transiently improves auditory function in a mouse model of recessive deafness. Sci. Transl. Med. 2020, 12. [Google Scholar] [CrossRef]
  130. György, B.; Nist-Lund, C.; Pan, B.; Asai, Y.; Karavitaki, K.D.; Kleinstiver, B.P.; Garcia, S.; Zaborowski, M.; Solanes, P.; Spataro, S.; et al. Allele-specific gene editing prevents deafness in a model of dominant progressive hearing loss. Nat. Med. 2019, 25, 1123–1130. [Google Scholar] [CrossRef]
  131. Hu, X.; Wang, J.; Yao, X.; Xiao, Q.; Xue, Y.; Wang, S.; Shi, L.; Shu, Y.; Li, H.; Yang, H. Screened AAV variants permit efficient transduction access to supporting cells and hair cells. Cell Discov. 2019, 5, e115–e119. [Google Scholar] [CrossRef]
  132. Zheng, Z.; Li, G.; Cui, C.; Wang, F.; Wang, X.; Xu, Z.; Guo, H.; Chen, Y.; Tang, H.; Wang, D.; et al. Preventing autosomal-dominant hearing loss in Bth mice with CRISPR/CasRx-based RNA editing. Signal Transduct. Target. Ther. 2022, 7, 1–13. [Google Scholar] [CrossRef]
  133. Noh, B.; Rim, J.H.; Gopalappa, R.; Lin, H.; Kim, K.M.; Kang, M.J.; Gee, H.Y.; Choi, J.Y.; Kim, H.H.; Jung, J. In vivo outer hair cell gene editing ameliorates progressive hearing loss in dominant-negative Kcnq4 murine model. Theranostics 2022, 12, 2465–2482. [Google Scholar] [CrossRef]
  134. Yin, H.; Xue, W.; Chen, S.; Bogorad, R.L.; Benedetti, E.; Grompe, M.; Koteliansky, V.; Sharp, P.A.; Jacks, T.E.; Anderson, D.G. Genome editing with Cas9 in adult mice corrects a disease mutation and phenotype. Nat. Biotechnol. 2014, 32, 551–553. [Google Scholar] [CrossRef]
  135. VanLith, M.C.; Guthman, M.R.; Nicolas, M.C.T.; Allen, M.K.; Du, Z.; Joo, D.J.; Nyberg, S.L.; Lillegard, J.B.; Hickey, R.D. Curative Ex Vivo Hepatocyte-Directed Gene Editing in a Mouse Model of Hereditary Tyrosinemia Type 1. Hum. Gene Ther. 2018, 29, 1315–1326. [Google Scholar] [CrossRef]
  136. Zhang, Q.-S.; Tiyaboonchai, A.; Nygaard, S.; Baradar, K.; Major, A.; Balaji, N.; Grompe, M. Induced Liver Regeneration Enhances CRISPR/Cas9-Mediated Gene Repair in Tyrosinemia Type 1. Hum. Gene Ther. 2021, 32, 294–301. [Google Scholar] [CrossRef]
  137. Bin Liu, B.; Dong, X.; Cheng, H.; Zheng, C.; Chen, Z.; Rodríguez, T.C.; Liang, S.-Q.; Xue, W.; Sontheimer, E.J. A split prime editor with untethered reverse transcriptase and circular RNA template. Nat. Biotechnol. 2022, 40, 1388–1393. [Google Scholar] [CrossRef]
  138. Yin, H.; Song, C.-Q.; Dorkin, J.R.; Zhu, L.J.; Li, Y.; Wu, Q.; Park, A.; Yang, J.; Suresh, S.; Bizhanova, A.; et al. Therapeutic genome editing by combined viral and non-viral delivery of CRISPR system components in vivo. Nat. Biotechnol. 2016, 34, 328–333. [Google Scholar] [CrossRef]
  139. Song, C.-Q.; Jiang, T.; Richter, M.; Rhym, L.H.; Koblan, L.W.; Zafra, M.P.; Schatoff, E.M.; Doman, J.L.; Cao, Y.; Dow, L.E.; et al. Adenine base editing in an adult mouse model of tyrosinaemia. Nat. Biomed. Eng. 2019, 4, 125–130. [Google Scholar] [CrossRef]
  140. Jiang, T.; Zhang, X.-O.; Weng, Z.; Xue, W. Deletion and replacement of long genomic sequences using prime editing. Nat. Biotechnol. 2021, 40, 227–234. [Google Scholar] [CrossRef]
  141. Chung, S.H.; Mollhoff, I.N.; Nguyen, U.; Nguyen, A.; Stucka, N.; Tieu, E.; Manna, S.; Meleppat, R.K.; Zhang, P.; Nguyen, E.L.; et al. Factors Impacting Efficacy of AAV-Mediated CRISPR-Based Genome Editing for Treatment of Choroidal Neovascularization. Mol. Ther.-Methods Clin. Dev. 2020, 17, 409–417. [Google Scholar] [CrossRef] [PubMed]
  142. Koo, T.; Park, S.W.; Jo, D.H.; Kim, D.; Kim, J.H.; Cho, H.-Y.; Kim, J.; Kim, J.H.; Kim, J.-S. CRISPR-LbCpf1 prevents choroidal neovascularization in a mouse model of age-related macular degeneration. Nat. Commun. 2018, 9, 1855. [Google Scholar] [CrossRef] [PubMed]
  143. Kim, K.; Park, S.W.; Kim, J.H.; Lee, S.H.; Kim, D.; Koo, T.; Kim, K.-E.; Kim, J.H.; Kim, J.S. Genome surgery using Cas9 ribonucleoproteins for the treatment of age-related macular degeneration. Genome Res. 2017, 27, 419–426. [Google Scholar] [CrossRef] [PubMed]
  144. Ling, S.; Yang, S.; Hu, X.; Yin, D.; Dai, Y.; Qian, X.; Wang, D.; Pan, X.; Hong, J.; Sun, X.; et al. Lentiviral delivery of co-packaged Cas9 mRNA and a Vegfa-targeting guide RNA prevents wet age-related macular degeneration in mice. Nat. Biomed. Eng. 2021, 5, 144–156. [Google Scholar] [CrossRef] [PubMed]
  145. Lee, J.-S.; Lee, J.Y.; Song, D.W.; Bae, H.S.; Doo, H.M.; Yu, H.S.; Lee, K.J.; Kim, H.K.; Hwang, H.; Kwak, G.; et al. Targeted PMP22 TATA-box editing by CRISPR/Cas9 reduces demyelinating neuropathy of Charcot-Marie-Tooth disease type 1A in mice. Nucleic Acids Res. 2019, 48, 130–140. [Google Scholar] [CrossRef]
  146. Jo, D.H.; Song, D.W.; Cho, C.S.; Kim, U.G.; Lee, K.J.; Lee, K.; Park, S.W.; Kim, D.; Kim, J.H.; Kim, J.S.; et al. CRISPR-Cas9–mediated therapeutic editing of Rpe65 ameliorates the disease phenotypes in a mouse model of Leber congenital amaurosis. Sci. Adv. 2019, 5, eaax1210. [Google Scholar] [CrossRef]
  147. Suh, S.; Choi, E.H.; Leinonen, H.; Foik, A.T.; Newby, G.A.; Yeh, W.-H.; Dong, Z.; Kiser, P.D.; Lyon, D.C.; Liu, D.R.; et al. Restoration of visual function in adult mice with an inherited retinal disease via adenine base editing. Nat. Biomed. Eng. 2020, 5, 169–178. [Google Scholar] [CrossRef]
  148. Choi, E.H.; Suh, S.; Foik, A.T.; Leinonen, H.; Newby, G.A.; Gao, X.D.; Banskota, S.; Hoang, T.; Du, S.W.; Dong, Z.; et al. In vivo base editing rescues cone photoreceptors in a mouse model of early-onset inherited retinal degeneration. Nat. Commun. 2022, 13, 1830. [Google Scholar] [CrossRef]
  149. Jang, H.; Jo, D.H.; Cho, C.S.; Shin, J.H.; Seo, J.H.; Yu, G.; Gopalappa, R.; Kim, D.; Cho, S.-R.; Kim, J.H.; et al. Application of prime editing to the correction of mutations and phenotypes in adult mice with liver and eye diseases. Nat. Biomed. Eng. 2021, 6, 181–194. [Google Scholar] [CrossRef]
  150. Park, H.; Oh, J.; Shim, G.; Cho, B.; Chang, Y.; Kim, S.; Baek, S.; Kim, H.; Shin, J.; Choi, H.; et al. In vivo neuronal gene editing via CRISPR–Cas9 amphiphilic nanocomplexes alleviates deficits in mouse models of Alzheimer’s disease. Nat. Neurosci. 2019, 22, 524–528. [Google Scholar] [CrossRef]
  151. Park, H.; Kim, J. Activation of melatonin receptor 1 by CRISPR-Cas9 activator ameliorates cognitive deficits in an Alzheimer’s disease mouse model. J. Pineal Res. 2022, 72, e12787. [Google Scholar] [CrossRef]
  152. Zhao, H.; Li, Y.; He, L.; Pu, W.; Yu, W.; Li, Y.; Wu, Y.-T.; Xu, C.; Wei, Y.; Ding, Q.; et al. In Vivo AAV-CRISPR/Cas9–Mediated Gene Editing Ameliorates Atherosclerosis in Familial Hypercholesterolemia. Circulation 2020, 141, 67–79. [Google Scholar] [CrossRef]
  153. Huang, J.Y.; Kan, S.-H.; Sandfeld, E.K.; Dalton, N.D.; Rangel, A.D.; Chan, Y.; Davis-Turak, J.; Neumann, J.; Wang, R.Y. CRISPR-Cas9 generated Pompe knock-in murine model exhibits early-onset hypertrophic cardiomyopathy and skeletal muscle weakness. Sci. Rep. 2020, 10, 1–13. [Google Scholar] [CrossRef]
  154. Kan, S.-h.; Huang, J.Y.; Harb, J.; Rha, A.; Dalton, N.D.; Christensen, C.; Chan, Y.; Davis-Turak, J.; Neumann, J.; Wang, R.Y. CRISPR-mediated generation and characterization of a Gaa homozygous c.1935C>A (p.D645E) Pompe disease knock-in mouse model recapitulating human infantile onset-Pompe disease. Sci. Rep. 2022, 12, 21576. [Google Scholar] [CrossRef]
  155. Santiago-Fernández, O.; Osorio, F.G.; Quesada, V.; Rodríguez, F.; Basso, S.; Maeso, D.; Rolas, L.; Barkaway, A.; Nourshargh, S.; Folgueras, A.R.; et al. Development of a CRISPR/Cas9-based therapy for Hutchinson–Gilford progeria syndrome. Nat. Med. 2019, 25, 423–426. [Google Scholar] [CrossRef]
  156. Beyret, E.; Liao, H.-K.; Yamamoto, M.; Hernandez-Benitez, R.; Fu, Y.; Erikson, G.; Reddy, P.; Belmonte, J.C.I. Single-dose CRISPR–Cas9 therapy extends lifespan of mice with Hutchinson–Gilford progeria syndrome. Nat. Med. 2019, 25, 419–422. [Google Scholar] [CrossRef]
  157. Berntorp, E.; Fischer, K.; Hart, D.P.; Mancuso, M.E.; Stephensen, D.; Shapiro, A.D.; Blanchette, V. Haemophilia. Nat. Rev. Dis. Prim. 2021, 7, e00278. [Google Scholar] [CrossRef]
  158. Hardiman, O.; Al-Chalabi, A.; Chio, A.; Corr, E.M.; Logroscino, G.; Robberecht, W.; Shaw, P.J.; Simmons, Z.; van den Berg, L.H. Amyotrophic lateral sclerosis. Nat. Rev. Dis. Prim. 2017, 3, 17071. [Google Scholar] [CrossRef]
  159. Tang, L.; Ma, Y.; Liu, X.-L.; Chen, L.; Fan, D.-S. Better survival in female SOD1-mutant patients with ALS: A study of SOD1-related natural history. Transl. Neurodegener. 2019, 8, 2. [Google Scholar] [CrossRef]
  160. Ballesteros, A.; Swartz, K.J. Regulation of membrane homeostasis by TMC1 mechanoelectrical transduction channels is essential for hearing. Sci. Adv. 2022, 8, e5550. [Google Scholar] [CrossRef]
  161. Zenner, H.P. The Role of Outer Hair Cell Damage in the Loss of Hearing. Ear Nose Throat J. 1997, 76, 140–144. [Google Scholar] [CrossRef] [PubMed]
  162. Chinsky, J.M.; Singh, R.; Ficicioglu, C.; van Karnebeek, C.D.M.; Grompe, M.; Mitchell, G.; Waisbren, S.E.; Gucsavas-Calikoglu, M.; Wasserstein, M.P.; Coakley, K.; et al. Diagnosis and treatment of tyrosinemia type I: A US and Canadian consensus group review and recommendations. Genet. Med. 2017, 19, 1380. [Google Scholar] [CrossRef] [PubMed]
  163. Fleckenstein, M.; Keenan, T.D.L.; Guymer, R.H.; Chakravarthy, U.; Schmitz-Valckenberg, S.; Klaver, C.C.; Wong, W.T.; Chew, E.Y. Age-related macular degeneration. Nat. Rev. Dis. Prim. 2021, 7, e265–e272. [Google Scholar] [CrossRef] [PubMed]
  164. Szigeti, K.; Lupski, J.R. Charcot–Marie–Tooth disease. Eur. J. Hum. Genet. 2009, 17, 703–710. [Google Scholar] [CrossRef] [PubMed]
  165. den Hollander, A.I.; Roepman, R.; Koenekoop, R.K.; Cremers, F.P. Leber congenital amaurosis: Genes, proteins and disease mechanisms. Prog. Retin. Eye Res. 2008, 27, 391–419. [Google Scholar] [CrossRef]
  166. Wang, J.; Gu, B.J.; Masters, C.L.; Wang, Y.-J. A systemic view of Alzheimer disease—Insights from amyloid-β metabolism beyond the brain. Nat. Rev. Neurol. 2017, 13, 612–623. [Google Scholar] [CrossRef]
  167. Shukla, M.; Govitrapong, P.; Boontem, P.; Reiter, R.J.; Satayavivad, J. Mechanisms of Melatonin in Alleviating Alzheimer’s Disease. Curr. Neuropharmacol. 2017, 15, 1010–1031. [Google Scholar] [CrossRef]
  168. Defesche, J.C.; Gidding, S.S.; Harada-Shiba, M.; Hegele, R.A.; Santos, R.D.; Wierzbicki, A.S. Familial hypercholesterolaemia. Nat. Rev. Dis. Prim. 2017, 3, 17093. [Google Scholar] [CrossRef]
  169. Civeira, F.F.; De Castro-Orós, I.; Pocoví, M. The genetic basis of familial hypercholesterolemia: Inheritance, linkage, and mutations. Appl. Clin. Genet. 2010, 3, 53–64. [Google Scholar] [CrossRef]
  170. Kishnani, P.S.; Steiner, R.D.; Bali, D.; Berger, K.; Byrne, B.J.; Case, L.E.; Crowley, J.F.; Downs, S.; Howell, R.R.; Kravitz, R.M.; et al. Pompe disease diagnosis and management guideline. Genet. Med. 2006, 8, 267–288. [Google Scholar] [CrossRef] [Green Version]
  171. Burtner, C.R.; Kennedy, B.K. Progeria syndromes and ageing: What is the connection? Nat. Rev. Mol. Cell Biol. 2010, 11, 567–578. [Google Scholar] [CrossRef]
  172. Eriksson, M.; Brown, W.T.; Gordon, L.B.; Glynn, M.W.; Singer, J.; Scott, L.; Erdos, M.R.; Robbins, C.M.; Moses, T.Y.; Berglund, P.; et al. Recurrent de novo point mutations in lamin A cause Hutchinson–Gilford progeria syndrome. Nature 2003, 423, 293–298. [Google Scholar] [CrossRef]
  173. Mullard, A. Gene-editing pipeline takes off. Nat. Rev. Drug Discov. 2020, 19, 367–372. [Google Scholar] [CrossRef]
  174. Schacker, M.; Seimetz, D. From fiction to science: Clinical potentials and regulatory considerations of gene editing. Clin. Transl. Med. 2019, 8, 1–16. [Google Scholar] [CrossRef]
  175. Wu, Z.; Zhang, Y.; Yu, H.; Pan, D.; Wang, Y.; Wang, Y.; Li, F.; Liu, C.; Nan, H.; Chen, W.; et al. Programmed genome editing by a miniature CRISPR-Cas12f nuclease. Nat. Chem. Biol. 2021, 17, 1132–1138. [Google Scholar] [CrossRef]
  176. Kim, D.Y.; Lee, J.M.; Bin Moon, S.; Chin, H.J.; Park, S.; Lim, Y.; Kim, D.; Koo, T.; Ko, J.-H.; Kim, Y.-S. Efficient CRISPR editing with a hypercompact Cas12f1 and engineered guide RNAs delivered by adeno-associated virus. Nat. Biotechnol. 2021, 40, 94–102. [Google Scholar] [CrossRef]
  177. Wang, Y.; Wang, Y.; Pan, D.; Yu, H.; Zhang, Y.; Chen, W.; Li, F.; Wu, Z.; Ji, Q. Guide RNA engineering enables efficient CRISPR editing with a miniature Syntrophomonas palmitatica Cas12f1 nuclease. Cell Rep. 2022, 40, 111418. [Google Scholar] [CrossRef]
  178. Sun, A.; Li, C.-P.; Chen, Z.; Zhang, S.; Li, D.-Y.; Yang, Y.; Li, L.-Q.; Zhao, Y.; Wang, K.; Li, Z.; et al. The compact Casπ (Cas12l) ‘bracelet’ provides a unique structural platform for DNA manipulation. Cell Res. 2023. [Google Scholar] [CrossRef]
  179. Zhang, S.; Song, L.; Yuan, B.; Zhang, C.; Cao, J.; Chen, J.; Qiu, J.; Tai, Y.; Chen, J.; Qiu, Z.; et al. TadA reprogramming to generate potent miniature base editors with high precision. Nat. Commun. 2023, 14, 1–12. [Google Scholar] [CrossRef]
  180. Qiu, M.; Glass, Z.; Chen, J.; Haas, M.; Jin, X.; Zhao, X.; Rui, X.; Ye, Z.; Li, Y.; Zhang, F.; et al. Lipid nanoparticle-mediated codelivery of Cas9 mRNA and single-guide RNA achieves liver-specific in vivo genome editing of Angptl3. Proc. Natl. Acad. Sci. USA 2021, 118, e2020401118. [Google Scholar] [CrossRef]
  181. Rurik, J.G.; Tombácz, I.; Yadegari, A.; Fernández, P.O.M.; Shewale, S.V.; Li, L.; Kimura, T.; Soliman, O.Y.; Papp, T.E.; Tam, Y.K.; et al. CAR T cells produced in vivo to treat cardiac injury. Science 2022, 375, 91–96. [Google Scholar] [CrossRef] [PubMed]
  182. Gillmore, J.D.; Gane, E.; Taubel, J.; Kao, J.; Fontana, M.; Maitland, M.L.; Seitzer, J.; O’Connell, D.; Walsh, K.R.; Wood, K.; et al. CRISPR-Cas9 In Vivo Gene Editing for Transthyretin Amyloidosis. N. Engl. J. Med. 2021, 385, 493–502. [Google Scholar] [CrossRef] [PubMed]
  183. Banskota, S.; Raguram, A.; Suh, S.; Du, S.W.; Davis, J.R.; Choi, E.H.; Wang, X.; Nielsen, S.C.; Newby, G.A.; Randolph, P.B.; et al. Engineered virus-like particles for efficient in vivo delivery of therapeutic proteins. Cell 2022, 185, 250–265.e16. [Google Scholar] [CrossRef] [PubMed]
  184. Kim, S.; Jeong, Y.K.; Cho, C.S.; Lee, S.; Sohn, C.H.; Kim, J.H.; Jeong, Y.; Jo, D.H.; Bae, S.; Lee, H. Enhancement of Gene Editing and Base Editing with Therapeutic Ribonucleoproteins through In Vivo Delivery Based on Absorptive Silica Nanoconstruct. Adv. Health Mater. 2022, 1002, 2201825. [Google Scholar] [CrossRef]
  185. Ohtsuka, M.; Sato, M.; Miura, H.; Takabayashi, S.; Matsuyama, M.; Koyano, T.; Arifin, N.; Nakamura, S.; Wada, K.; Gurumurthy, C.B. i-GONAD: A robust method for in situ germline genome engineering using CRISPR nucleases. Genome Biol. 2018, 19, 25. [Google Scholar] [CrossRef]
  186. Imai, Y.; Tanave, A.; Matsuyama, M.; Koide, T. Efficient genome editing in wild strains of mice using the i-GONAD method. Sci. Rep. 2022, 12, 13821. [Google Scholar] [CrossRef]
  187. Shang, R.; Zhang, H.; Bi, P. Generation of mouse conditional knockout alleles in one step using the i-GONAD method. Genome Res. 2020, 31, 121–130. [Google Scholar] [CrossRef]
  188. Quadros, R.M.; Miura, H.; Harms, D.W.; Akatsuka, H.; Sato, T.; Aida, T.; Redder, R.; Richardson, G.P.; Inagaki, Y.; Sakai, D.; et al. Easi-CRISPR: A robust method for one-step generation of mice carrying conditional and insertion alleles using long ssDNA donors and CRISPR ribonucleoproteins. Genome Biol. 2017, 18, 92. [Google Scholar] [CrossRef]
  189. Miura, H.; Quadros, R.M.; Gurumurthy, C.B.; Ohtsuka, M. Easi-CRISPR for creating knock-in and conditional knockout mouse models using long ssDNA donors. Nat. Protoc. 2017, 13, 195–215. [Google Scholar] [CrossRef]
  190. Gu, B.; Posfai, E.; Rossant, J. Efficient generation of targeted large insertions by microinjection into two-cell-stage mouse embryos. Nat. Biotechnol. 2018, 36, 632–637. [Google Scholar] [CrossRef]
  191. Gu, B.; Posfai, E.; Gertsenstein, M.; Rossant, J. Efficient Generation of Large-Fragment Knock-In Mouse Models Using 2-Cell (2C)-Homologous Recombination (HR)-CRISPR. Curr. Protoc. Mouse Biol. 2020, 10, e67. [Google Scholar] [CrossRef]
  192. Chen, S.; Sun, S.; Moonen, D.; Lee, C.; Lee, A.Y.-F.; Schaffer, D.V.; He, L. CRISPR-READI: Efficient Generation of Knockin Mice by CRISPR RNP Electroporation and AAV Donor Infection. Cell Rep. 2019, 27, 3780–3789.e4. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Delivery strategies for gene editing using programmable nucleases in mice. The genome editors, including ZFNs, TALENs, CRISPR/Cas, PE, CBE, and ABE, can be delivered to mice using viral (AAVs and lentiviruses) or nonviral particles (SN, nanoparticles, and eVLPs) via microinjections or electroporation. In addition, the programmable nucleases can be directly delivered to mice using the i-GONAD method without any ex vivo modifications. ZFNs, zinc finger nucleases; TALENs, transcription activator-like effector nucleases; CRISPR, clustered regularly interspaced short palindromic repeats; Cas, CRISPR-associated protein; PE, prime editor; CBE, cytosine base editor; ABE, adenine base editor; AAVs, adeno-associated viral systems, SN, sponge-like silica nanoconstruct; eVLPs, engineered DNA-free virus-like particles; i-GONAD, improved genome editing via oviductal nucleic acid delivery. * This illustration was created with BioRender.com (accessed on 1 February 2023).
Figure 1. Delivery strategies for gene editing using programmable nucleases in mice. The genome editors, including ZFNs, TALENs, CRISPR/Cas, PE, CBE, and ABE, can be delivered to mice using viral (AAVs and lentiviruses) or nonviral particles (SN, nanoparticles, and eVLPs) via microinjections or electroporation. In addition, the programmable nucleases can be directly delivered to mice using the i-GONAD method without any ex vivo modifications. ZFNs, zinc finger nucleases; TALENs, transcription activator-like effector nucleases; CRISPR, clustered regularly interspaced short palindromic repeats; Cas, CRISPR-associated protein; PE, prime editor; CBE, cytosine base editor; ABE, adenine base editor; AAVs, adeno-associated viral systems, SN, sponge-like silica nanoconstruct; eVLPs, engineered DNA-free virus-like particles; i-GONAD, improved genome editing via oviductal nucleic acid delivery. * This illustration was created with BioRender.com (accessed on 1 February 2023).
Genes 14 00483 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Phan, H.T.L.; Kim, K.; Lee, H.; Seong, J.K. Progress in and Prospects of Genome Editing Tools for Human Disease Model Development and Therapeutic Applications. Genes 2023, 14, 483. https://doi.org/10.3390/genes14020483

AMA Style

Phan HTL, Kim K, Lee H, Seong JK. Progress in and Prospects of Genome Editing Tools for Human Disease Model Development and Therapeutic Applications. Genes. 2023; 14(2):483. https://doi.org/10.3390/genes14020483

Chicago/Turabian Style

Phan, Hong Thi Lam, Kyoungmi Kim, Ho Lee, and Je Kyung Seong. 2023. "Progress in and Prospects of Genome Editing Tools for Human Disease Model Development and Therapeutic Applications" Genes 14, no. 2: 483. https://doi.org/10.3390/genes14020483

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

Article Metrics

Back to TopTop