Abstract

Genomic editing to correct disease-causing mutations is a promising approach for the treatment of human diseases. As a simple and programmable nuclease-based genomic editing tool, the clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein 9 (Cas9) system has substantially improved the ability to make precise changes in the human genome. Rapid development of CRISPR-based technologies in recent years has expanded its application scope and promoted CRISPR-based therapies in preclinical trails. Here, we review the application of the CRISPR system over the last 2 years; including its development and application in base editing, transcription modulation and epigenetic editing, genomic-scale screening, and cell and embryo therapy. Finally, the prospects and challenges related to application of CRISPR/Cas9 technologies are discussed.

Introduction

The clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein (Cas) system is a bacterial and archaebacterial defense mechanism against phage infection and plasmid transfer in nature. In recent years, this system has garnered increasing attention as an effective and simple genome-engineering tool and has revolutionized the life sciences. It has been recognized for its potentially transformative applications in transcriptional perturbation, epigenetic modulation, base editing, high-throughput genetic screening and generation of animal or cell models of diseases (1).

Based on the effector protein organization, CRISPR/Cas system is classified into two distinct classes that are subdivided into 6 types and 19 subtypes. Class 1 CRISPR/Cas systems utilize multi-protein effector complexes, whereas class 2 CRISPR/Cas systems utilize single-protein effectors (2,3). The diverse CRISPR/Cas systems have a number of features in common, such as the use of short DNA sequences known as ‘spacers’ to direct the targeting of Cas proteins. In addition, there is a requirement for a conserved sequence to aid targeting, called the protospacer adjacent motif (PAM) for DNA-targeted Cas proteins, or the protospacer flanking sequence (PFS) for RNA-targeted Cas proteins (4). The Cas9 endonuclease from Streptococcus pyogenes, which belong to the class 2 CRISPR system, has been the most widely used CRISPR system for gene editing. Currently, there are ongoing clinical trials using CRISPR/Cas9-edited human cells, for example, to treat cancer. In the following sections, we review the applications of CRISPR/Cas9 technologies in the last 2 years; including base editing, transcription modulation and epigenetic editing, genomic-scale screening, and cell and embryo therapy (Fig. 1). We will also discuss the limitations as well as regulatory and ethical implications of this transformative set of technologies.

Applications of CRISPR/Cas9 technology. (A) Traditional genome editing: CRISPR/Cas9 is used for making DSB at particular sites. These introduced DSBs are mended by either NHEJ or HDR. (B) Base editing: A base editor consists of a catalytically modified Cas9, deoxyadenosine deaminase enzymes and TadA fusion proteins complex. Through deoxyadenosine-deaminase action, adenosine is transformed into inosine. (C) Transcription modulation: CRISPR transcription modulation system has been developed by modifying dCas9 protein. Combined with activate or repression domain (such as VP64 and KRAB), dCas9 can activate or repress, respectively, DNA transcription without changing DNA sequence. (D) Epigenetic modulation: DNA methylation and histone modification; epigenetic effectors merged with dCas9 could change epigenetic states of target loci, such as through DNA methylation or histone modifications. (E) Pooled CRISPR screen strategy. In pooled CRISPR screens, gRNAs are synthesized, cloned and constructed as a pool. Pooled screens have dissected genetic networks with the help of NGS like single-cell RNA-seq.
Figure 1.

Applications of CRISPR/Cas9 technology. (A) Traditional genome editing: CRISPR/Cas9 is used for making DSB at particular sites. These introduced DSBs are mended by either NHEJ or HDR. (B) Base editing: A base editor consists of a catalytically modified Cas9, deoxyadenosine deaminase enzymes and TadA fusion proteins complex. Through deoxyadenosine-deaminase action, adenosine is transformed into inosine. (C) Transcription modulation: CRISPR transcription modulation system has been developed by modifying dCas9 protein. Combined with activate or repression domain (such as VP64 and KRAB), dCas9 can activate or repress, respectively, DNA transcription without changing DNA sequence. (D) Epigenetic modulation: DNA methylation and histone modification; epigenetic effectors merged with dCas9 could change epigenetic states of target loci, such as through DNA methylation or histone modifications. (E) Pooled CRISPR screen strategy. In pooled CRISPR screens, gRNAs are synthesized, cloned and constructed as a pool. Pooled screens have dissected genetic networks with the help of NGS like single-cell RNA-seq.

Base Editing

Traditional CRISPR/Cas9 protein-RNA complexes localize to a target DNA sequence through base pairing with a guide RNA, and natively create a dsDNA break (DSB) at the locus specified by the guide RNA (1). In response to DSBs, cellular DNA repair processes result in random insertions or deletions at the site of DNA cleavage through homology-directed repair (HDR) or non-homologous end joining (NHEJ), depending on the presence or absence of a homologous DNA template, respectively. However, HDR-dependent genome editing is limited by low efficiency arising from competition with NHEJ outcomes and from the dependence of HDR on mitosis. Therefore, the primary strategies of modifying eukaryotic genomes using HDR are more unpredictable than the prospection of precise mutation correction (5,6). While researchers have attempted to increase the efficiency of HDR by suppressing NHEJ, the strategies for correction of point mutations using HDR under therapeutically relevant conditions remain inefficient (7,8). In this instance, the third-generation base editor (BE3), a single protein, was designed to allow the direct and stable transformation of target DNA bases into an alternative in a programmable way, without DNA double-strand cleavage or a donor template (9). BE3 contains a tripartite fusion between Rattus norvegicus APOBEC1 cytidine deaminase, S. pyogenes Cas9n (D10A) and Bacillus subtilis bacteriophage PBS2 UGI. BE3 results in a higher base editing efficiency than HDR, and a much lower indel frequency than nuclease-mediated approaches. Moreover, multiple studies have shown that BE3 has fewer off-target editing events than Cas9 (10,11).

Subsequently, BE4 was designed based on B3 to increase the efficiency of C: G to T: A base editing by approximately 50%, while halving the frequency of undesired by-products with respect to those generated by BE3. Fusion of this optimized construct to the DSB binding protein Gam from bacteriophage Mu yielded BE4-Gam, a base editor that retains the optimized properties of BE4 while further reducing indel formation (12). DNA glycosylase inhibitor protein (e.g. UGI) was also used to enhance base editing efficiency. In addition to BE3 and BE4, other cytidine deaminase enzymes like activation-induced cytidine deaminase (AID) have also been developed to address the inherent limitations of using DSBs for gene correction (termed ‘CRISPR-X’). Target-AID is a BE3-like base editor, and is a synthetic complex of dCas9 and PmCDA1. Target-AID results in the base substitution of C or G into T or A with a shifted deamination window compared with BE3 (13). Genome-wide specificity of base editing approaches continues to be evaluated and improved (10).

Over the last 2 years, base editing tools and related technologies have been successfully used by many researchers in a wide range of applications, including plant genome editing, in vivo mammalian genome editing, targeted mutagenesis and knockout studies (13–21). The ability to create or remove a single-nucleotide variation (SNV) in target genes makes BEs valuable tools in genetic, therapeutic and agricultural applications. BEs can also create nonsense mutations by generating premature TGA (opal), TAG (amber) or TAA (ochre) stop codons when targeted to CGA (Arg), CAG (Gln) and CAA (Gln) (15,16,22). This method has the advantage over traditional Cas9 in gene knockout experiments, as Cas9 often induces in-frame indels in a protein-coding gene, which can still produce functional proteins, and frame-shifting indels, leading to translation of out-of-frame polypeptide sequences that can be immunogenic and may have unknown effects in cells (23). Base editing can also be combined with a library of sgRNAs targeted to a gene of interest and used for high-throughput screening of gain-of-function variants in cells (14,18).

Transcription Modulation and Epigenetic Editing

The dCas9 protein is a Cas9 variant which is capable of binding to the target sequence but unable to cleave its target (24). This protein has been adopted as a DNA-binding platform for transcription modulation and epigenetic editing, and engineered by using a variety of effector domains.

In the previous studies, dCas9 use focused on target gene activation (TGA) by fusion with conventional transcriptional activators such as VP64, p65 or a subunit of RNA polymerase (25–28). However, the dCas9-VP64 system was not very effective and was also unable to stimulate robust TGA by using a single-guide RNA (sgRNA) (29–31). Thus, fusion or recruitment of multiple transcriptional activation domains to the dCas9/gRNA complex (32), synergistic activation mediator (SAM) (33,34) or dCas9-Suntag (35,36) were performed to improve the activation capacity and expand the range of applications. These systems could induce very high activation by using several sgRNAs, or even one sgRNA for each target gene, enabling high levels of activation in vitro (37–39). However, in vivo use remains a challenge (1,40), possibly attributed to the lower efficiency of in vivo transferring of dCas9 fusion proteins and the lower level of TGA in vivo. In addition, sequences encoding the dCas9/gRNA and co-transcriptional activator complexes exceed the capacity of most common viral vectors like AAV, which is the most promising vector for gene delivery in vivo (41). In a recent report, an in vivo CRISPR/Cas9 TGA system was created to overcome this problem. In this system, the transcriptional activators were separated from dCas9 to shorten the coding sequence. Other researchers tried to optimize the shortening of gRNAs [14 or 15 base pairs (bp) rather than 20 bp] containing motifs that tether bacteriophage MS2 coat protein domains fused to the MS2: P65: HSF1 (MPH) transcriptional activation complex; this system achieved high efficiency TGA in organs of living animals (41).

dCas9 can also block target gene transcription by fusing a repressive effector domain such as the Krüppel-associated box (KRAB) (27,42). KRAB repression is mediated by repressive histone modifications such as H3K9me3. By utilizing epigenome-modifying repressors, including Lys-specific histone demethylase 1 (LSD1) (43), histone deacetylase (HDAC) (44), DNA methyltransferases DNMT3A and MQ1 (39,45,46), and mSin3 interaction domains (47), the scope of applying CRISPR repression has been extended to epigenetic editing. Similarly, epigenome editing approaches can also be used for targeted transcriptional activation, such as dCas9 fused with a DNA demethylase or a histone acetyltransferase (39,48). Recently, Klann et al. (48) developed a CRISPR/Cas9-based epigenomic regulatory element screening (CERES) system, which combines dCas9-p300Core with dCas9-KRAB to obtain both gain and loss of function information by targeting the same regions with a repressor and an activator.

Genomic-Scale Screening

In recent years, tools that use the RNA interference (RNAi) pathway, specifically short-hairpin RNAs (shRNAs), to perturb transcript levels have revolutionized screening approaches (49–51). However, this approach was limited by incompletely abrogate gene expression and high off-target effects, resulting in false positive (52,53). Some studies have shown that CRISPR can be adapted for genome-scale screening by combining Cas9 with pooled-guide RNA libraries and next-generation sequencing (NGS) (54). To date, two applications of CRISPR/Cas9-mediated genomic modifiers have been available for genome-wide screening, either by targeting about 20 000 genes or studying one specific signaling pathway or gene function (55). The general approach for screening of CRISPR is through loss-of-function screens, which is based on indel-prone NHEJ repair or repression of the underlying sequences (CRISPRi). The gain-of-function screen, which uses endogenous HDR or CRISPR activation (CRISPRa) approaches, is also employed for some specific applications (48). In addition, developments in the CRISPR system for RNA targeting or base editing will further broaden the research field of CRISPR screens (56,57).

Given the ever-growing CRISPR toolbox, a variety of high-throughput pooled screening options have become available for genome mutagenesis, transcription and epigenome modification, and base editing. These screens utilize pooled sgRNA libraries generated by cloning chip-synthesized oligonucleotides, which cover the entire human or mouse transcriptome, into lentiviral vectors. Upon transduction of the libraries into cells, sgRNAs inducing a selectable phenotype can be identified by NGS. Traditionally, CRISPR/Cas9 screens have been used to study intracellular phenotypes by combining with positive, negative or marker/reporter gene selection (58). Recently, single-cell RNAseq has been developed for the pooled CRISPR screen readout as an alternative strategy (‘Perturb-seq’), allowing capture of multi-dimensional transcriptional phenotypes to correlate with their respective genetic manipulation (59–61).

CRISPR/Cas9 library screening enables the identification of critical components in a variety of biological processes. Apart from being widely used to identify essential genes in mammalian cell lines (62), the CRISPR screen has recently been used to analyse synthetic lethal genes (63), HIV host dependency factors (64), genetic vulnerabilities in cancer (65,66), novel targets of diseases (67–69) and resistance phenotypes (70). Moreover, the identification of non-coding sequences and characterized enhancer elements and regulatory sequences belongs to a novel application scope of CRISPR screens. With tiling sgRNAs (‘saturating mutagenesis’), researchers could identify non-coding regions involved in resistance to pharmacological inhibitors (71), enhancers required for oncogene-induced senescence (72), or novel enhancer elements (71). CRISPR/Cas9 screens have also been extended to the identification of short-length miRNAs through combining CRISPR/dCas9 with a split-HRP-based reporter system (73). In addition, high-throughput genetic studies combined with CRISPR screens have also been performed in vivo by delivering gRNA and nuclease with vectors or using constitutive or inducible Cas9-expressing mouse models that only require gRNA delivery (65,74,75).

Applications in Cell Therapy

The CRISPR/Cas9 system has produced a revolution in the field of cell therapy, mainly involving immune cell therapy and stem cell therapy [42]. Treatment using ex vivo gene-editing T cells from patients with cancer or autoimmune diseases have shown promising results (28,76–78); one example being next-generation chimeric antigen receptor (CAR) T cells. The feasibility of using CAR T cells for targeted therapy of malignancies has been established through using electroporation of Cas9 ribonucleoproteins (RNPs) to transfect activated various targets including CXCR4, CCR5, PD-1 and CD7 in human T cells (54,76,79–82). The first clinical trials using CRISPR/Cas9-RGNs to create a PD-1 knockout in T cells have been approved for the treatment of muscle-invasive bladder cancer, castration-resistant prostate cancer, metastatic renal cancer and metastatic non-small cell lung cancer. These phase I clinical trials were commenced in 2016 (83). However, severe side effects due to cytokine release syndrome (CRS), neurotoxicity or on-target off-tumor toxicity are major obstacles for effective treatment of patients. To relieve such limitations related to T cell–based immunotherapies, CRISPR/Cas9 system was used to delete endogenous TCRs and HLA class I to generate the universal allogenic ‘off-the-shelf’ CAR T cells, or disrupt inhibitory receptors, such as CTLA-4 or PD-1 (76,79,84,85), or target CAR constructs to the endogenous TCR α constant locus in the primary human T cells with gene editing tools (86). Immune cell therapy will achieve greater breakthroughs in clinical application, as the US Food and Drug Administration (FDA) has approved two global CAR-T cell products targeting CD19 for the treatment of acute lymphoblastic leukemia and non-Hodgkin lymphoma in the past year.

Apart from T cells, ex vivo editing of induced pluripotent stem cells (iPSCs) and hematopoietic stem cells (HSCs) derived from the patient’s somatic cells also likely offer the opportunity for investigating the pathophysiological mechanisms of hereditary diseases. With this technology, researchers have made advances in the treatment of β-hemoglobinopathies, hemophilia B, hearing loss, Alzheimer disease, oculopathy, cardiac diseases and other genetic diseases (Table 1). Notably, the production of organoids from engineered iPSCs has developed rapidly in recent years. Patient-specific organoids offer unprecedented opportunities for studying phenotypes manifested at the cellular level, and provides a framework for both disease modeling and regenerative medicine based on the synthetic reconstitution of tissues with physiologically relevant structural and functional features that could be transplanted into patients (87). With these organoids, the in-depth mechanisms of muscular development and disease (116), epithelial development (117) and some other genetic diseases (118,119) have been revealed. A clinical trial for age-related macular degeneration (AMD) was performed whereby autologous iPSC-derived retinal pigment epithelial (RPE) cells were transplanted into the patient’s pathological tissue. Although this treatment could improve a patient’s symptoms and avoid immune rejection, the non-ideal therapeutic effects and tumorigenicity of iPSCs are still serious disadvantages (120).

Table 1.

Ex vivo editing of induced pluripotent stem cells (iPSCs) related to genetic diseases

Associated diseasesMutant genesCitation
Sandhoff diseaseExosaminidase subunit beta (HEXB)(88)
Long QT syndrome (LQTS)Calmodulin 2(CALM2)(89)
Duchenne muscular dystrophyX-linked dystrophin gene (DMD)(130)
Cleidocranial dysplasia (CCD)Runt-related transcription factor 2 (RUNX2)(90)
N370S GBA1 Parkinson’s disease (PD)Glucosylceramidase beta1 (GBA1)(91)
Coenzyme Q10 deficiencyCoenzyme Q4(COQ4)(92)
Danon diseaseLysosomal-associated membrane protein 2 (LAMP-2)(93)
Glanzmann thrombastheniaIntegrin subunit alpha 2b (ITGA2B)(94)
Metachromatic leukodystrophyArylsulfatase A (ARSA)(95)
Neuronal ceroid lipofuscinoses (Batten disease)CLN5, intracellular trafficking protein (CLN5)(96)
Niemann-Pick disease, types A and C1Sphingomyelin phosphodiesterase 1 (SMPD1), NPC intracellular cholesterol transporter 1 (NPC1)(97)
Pelizaeus-Merzbacher diseaseProteolipid protein 1 (PLP1)(98)
Pompe diseaseGlucosidase alpha, acid (GAA)(99)
Prader-Willi syndrome15q11.2-q13; Various; Chr.7(100,101)
Retinitis pigmentosaMER proto-oncogene, tyrosine kinase (MERTK)(102)
Smith-Lemli-Opitz syndrome7-Dehydrocholesterol reductase (DHCR7)(103)
Wolman disease (lysosomal acid lipase disease)Lipase A, lysosomal acid type (LIPA)(104)
Congenital neutropenia (SCN, Kostmann disease)HCLS1-associated protein X-1 (HAX1)(105)
X-linked RP (XLRP)Retinitis pigmentosa GTPase regulator (RPGR); nuclear receptor subfamily 2 group E member 3(Nr2e3)(106,117)
Leber congenital amaurosis (LCA)Centrosomal protein 290 [Homo sapiens (CEP290)](107)
MEN2ARet proto-oncogene (RET)(108)
Familial platelet disorder (FPD)Runt-related transcription factor 2 (Runx1)(109)
Parkinson's disease (PD)Triplication of the α-synuclein (SNCA)(110)
ADPresenilin (PSEN1)(111)
Hemophilia BCoagulation factor IX (F IX)(112)
Hereditary sensual deafnessMyosin VIIA (MYO7A); myosin XVA (MYO15A)(113,114)
Retinitis pigmentosaMEN1(115)
Associated diseasesMutant genesCitation
Sandhoff diseaseExosaminidase subunit beta (HEXB)(88)
Long QT syndrome (LQTS)Calmodulin 2(CALM2)(89)
Duchenne muscular dystrophyX-linked dystrophin gene (DMD)(130)
Cleidocranial dysplasia (CCD)Runt-related transcription factor 2 (RUNX2)(90)
N370S GBA1 Parkinson’s disease (PD)Glucosylceramidase beta1 (GBA1)(91)
Coenzyme Q10 deficiencyCoenzyme Q4(COQ4)(92)
Danon diseaseLysosomal-associated membrane protein 2 (LAMP-2)(93)
Glanzmann thrombastheniaIntegrin subunit alpha 2b (ITGA2B)(94)
Metachromatic leukodystrophyArylsulfatase A (ARSA)(95)
Neuronal ceroid lipofuscinoses (Batten disease)CLN5, intracellular trafficking protein (CLN5)(96)
Niemann-Pick disease, types A and C1Sphingomyelin phosphodiesterase 1 (SMPD1), NPC intracellular cholesterol transporter 1 (NPC1)(97)
Pelizaeus-Merzbacher diseaseProteolipid protein 1 (PLP1)(98)
Pompe diseaseGlucosidase alpha, acid (GAA)(99)
Prader-Willi syndrome15q11.2-q13; Various; Chr.7(100,101)
Retinitis pigmentosaMER proto-oncogene, tyrosine kinase (MERTK)(102)
Smith-Lemli-Opitz syndrome7-Dehydrocholesterol reductase (DHCR7)(103)
Wolman disease (lysosomal acid lipase disease)Lipase A, lysosomal acid type (LIPA)(104)
Congenital neutropenia (SCN, Kostmann disease)HCLS1-associated protein X-1 (HAX1)(105)
X-linked RP (XLRP)Retinitis pigmentosa GTPase regulator (RPGR); nuclear receptor subfamily 2 group E member 3(Nr2e3)(106,117)
Leber congenital amaurosis (LCA)Centrosomal protein 290 [Homo sapiens (CEP290)](107)
MEN2ARet proto-oncogene (RET)(108)
Familial platelet disorder (FPD)Runt-related transcription factor 2 (Runx1)(109)
Parkinson's disease (PD)Triplication of the α-synuclein (SNCA)(110)
ADPresenilin (PSEN1)(111)
Hemophilia BCoagulation factor IX (F IX)(112)
Hereditary sensual deafnessMyosin VIIA (MYO7A); myosin XVA (MYO15A)(113,114)
Retinitis pigmentosaMEN1(115)
Table 1.

Ex vivo editing of induced pluripotent stem cells (iPSCs) related to genetic diseases

Associated diseasesMutant genesCitation
Sandhoff diseaseExosaminidase subunit beta (HEXB)(88)
Long QT syndrome (LQTS)Calmodulin 2(CALM2)(89)
Duchenne muscular dystrophyX-linked dystrophin gene (DMD)(130)
Cleidocranial dysplasia (CCD)Runt-related transcription factor 2 (RUNX2)(90)
N370S GBA1 Parkinson’s disease (PD)Glucosylceramidase beta1 (GBA1)(91)
Coenzyme Q10 deficiencyCoenzyme Q4(COQ4)(92)
Danon diseaseLysosomal-associated membrane protein 2 (LAMP-2)(93)
Glanzmann thrombastheniaIntegrin subunit alpha 2b (ITGA2B)(94)
Metachromatic leukodystrophyArylsulfatase A (ARSA)(95)
Neuronal ceroid lipofuscinoses (Batten disease)CLN5, intracellular trafficking protein (CLN5)(96)
Niemann-Pick disease, types A and C1Sphingomyelin phosphodiesterase 1 (SMPD1), NPC intracellular cholesterol transporter 1 (NPC1)(97)
Pelizaeus-Merzbacher diseaseProteolipid protein 1 (PLP1)(98)
Pompe diseaseGlucosidase alpha, acid (GAA)(99)
Prader-Willi syndrome15q11.2-q13; Various; Chr.7(100,101)
Retinitis pigmentosaMER proto-oncogene, tyrosine kinase (MERTK)(102)
Smith-Lemli-Opitz syndrome7-Dehydrocholesterol reductase (DHCR7)(103)
Wolman disease (lysosomal acid lipase disease)Lipase A, lysosomal acid type (LIPA)(104)
Congenital neutropenia (SCN, Kostmann disease)HCLS1-associated protein X-1 (HAX1)(105)
X-linked RP (XLRP)Retinitis pigmentosa GTPase regulator (RPGR); nuclear receptor subfamily 2 group E member 3(Nr2e3)(106,117)
Leber congenital amaurosis (LCA)Centrosomal protein 290 [Homo sapiens (CEP290)](107)
MEN2ARet proto-oncogene (RET)(108)
Familial platelet disorder (FPD)Runt-related transcription factor 2 (Runx1)(109)
Parkinson's disease (PD)Triplication of the α-synuclein (SNCA)(110)
ADPresenilin (PSEN1)(111)
Hemophilia BCoagulation factor IX (F IX)(112)
Hereditary sensual deafnessMyosin VIIA (MYO7A); myosin XVA (MYO15A)(113,114)
Retinitis pigmentosaMEN1(115)
Associated diseasesMutant genesCitation
Sandhoff diseaseExosaminidase subunit beta (HEXB)(88)
Long QT syndrome (LQTS)Calmodulin 2(CALM2)(89)
Duchenne muscular dystrophyX-linked dystrophin gene (DMD)(130)
Cleidocranial dysplasia (CCD)Runt-related transcription factor 2 (RUNX2)(90)
N370S GBA1 Parkinson’s disease (PD)Glucosylceramidase beta1 (GBA1)(91)
Coenzyme Q10 deficiencyCoenzyme Q4(COQ4)(92)
Danon diseaseLysosomal-associated membrane protein 2 (LAMP-2)(93)
Glanzmann thrombastheniaIntegrin subunit alpha 2b (ITGA2B)(94)
Metachromatic leukodystrophyArylsulfatase A (ARSA)(95)
Neuronal ceroid lipofuscinoses (Batten disease)CLN5, intracellular trafficking protein (CLN5)(96)
Niemann-Pick disease, types A and C1Sphingomyelin phosphodiesterase 1 (SMPD1), NPC intracellular cholesterol transporter 1 (NPC1)(97)
Pelizaeus-Merzbacher diseaseProteolipid protein 1 (PLP1)(98)
Pompe diseaseGlucosidase alpha, acid (GAA)(99)
Prader-Willi syndrome15q11.2-q13; Various; Chr.7(100,101)
Retinitis pigmentosaMER proto-oncogene, tyrosine kinase (MERTK)(102)
Smith-Lemli-Opitz syndrome7-Dehydrocholesterol reductase (DHCR7)(103)
Wolman disease (lysosomal acid lipase disease)Lipase A, lysosomal acid type (LIPA)(104)
Congenital neutropenia (SCN, Kostmann disease)HCLS1-associated protein X-1 (HAX1)(105)
X-linked RP (XLRP)Retinitis pigmentosa GTPase regulator (RPGR); nuclear receptor subfamily 2 group E member 3(Nr2e3)(106,117)
Leber congenital amaurosis (LCA)Centrosomal protein 290 [Homo sapiens (CEP290)](107)
MEN2ARet proto-oncogene (RET)(108)
Familial platelet disorder (FPD)Runt-related transcription factor 2 (Runx1)(109)
Parkinson's disease (PD)Triplication of the α-synuclein (SNCA)(110)
ADPresenilin (PSEN1)(111)
Hemophilia BCoagulation factor IX (F IX)(112)
Hereditary sensual deafnessMyosin VIIA (MYO7A); myosin XVA (MYO15A)(113,114)
Retinitis pigmentosaMEN1(115)

Applications in the Modification of Human Embryo

Gene editing in human zygotes or embryos were previously unthinkable due to the low efficiency of traditional biological technologies. With the rapid development of CRISPR technology, researchers have attempted to edit human embryos using CRISPR and made some progress. Several scientific groups in China have reported genes corrected with the CRISPR/Cas9 technique in human zygotes or 3PN embryos (121,122). Recently, Ma et al. (123) described the correction of a pathogenic gene mutation in human embryos using CRISPR technology. They corrected the heterozygous MYBPC3 mutation in human pre-implantation embryos with precise CRISPR/Cas9-based targeting accuracy and high HDR efficiency by activating an endogenous, germline-specific DNA repair response. Other studies were the first to use the base editor system to correct HBB (A>G) mutation within the human embryonic genome (124,125). These breakthroughs highlight the tremendous potential of correcting homozygous and compound heterozygous mutations by base editing in human embryos.

Although the gene targeting efficiencies in the above studies were quite low and were accompanied by significant off-target effects as well as mosaicism, CRISPR/Cas systems have the potential of correcting heritable mutations in human embryos. Ma et al. (123) reported that 72.4% of human embryos fertilized with sperm carrying the heterozygous MYBPC3 mutation could carry two copies of the non-mutated gene copy after Cas9-sgRNA proteins injected, as opposed to 47.4% in untreated embryos. Moreover, only one mosaic embryo was detected, with results showing 100% targeting efficiency based on the absence of detected heterozygous blastomeres (123,126). Notably, embryonic genome editing with a base editor showed higher efficiency: targeted deep sequencing on injected embryos revealed that 17 out of 17 (100%) or 6 out of 9 (67%) embryos carried the targeted point mutations at the target site in FANCF or DNMT3B gene, respectively (125). This supports the supposition that base editors can be beneficial for correcting genetic defects in human embryos.

In addition, the above achievements have led to the debate of human embryo genome editing regarding ethical implications, especially clinical applications (127). Although some institutions like the National Academy of Sciences, Engineering, and Medicine in the United States have started the support for modifying human embryos by correcting mutations that cause serious disease when no reasonable alternatives exist (128), further progress is necessary before CRISPR/Cas9 techniques can become a reasonable and safe approach for clinical applications, especially in relation to embryo genome editing.

Limitations and Prospects

Despite being effectively useful in genome modification and functional regulation in biological systems, CRISPR/Cas9 systems have also encountered various inherent challenges. Among them, the off-target effect is the foremost challenge. In the last few years, researchers have attempted to make improvements by optimizing sgRNA, or artificially modifying the protein structure of Cas9, such as the PAM structure (129). In addition, there are investigations into anti-CRISPR proteins which promise to be a useful Cas9 off-switch for avoiding undesirable off-target effects by limiting the amount of time that Cas9 is active in the nucleus (130–132).

The delivery efficiency of Cas9 into cells or tissue is another problem that is preventing successful CRISPR-mediated genome editing. Initial CRISPR-mediated genome editing approaches relied on delivering plasmids or viral vectors encoding Cas9 and sgRNAs. Adeno-associated virus (AAV) vectors, which can facilitate gene transfer and episomal expression in non-dividing cells, are the mostly used vectors in Cas9 delivery. However, its application suffers from limitations associated with the size of encapsulated transgenes, pre-existing immunity against AAV vectors and CD8+ T cell-mediated adaptive response against the AAV capsid (133). Several studies have shown that the delivery format of active as9 protein/gRNA RNP complex through microinjection, liposome-mediated transfection, electroporation or nucleofection (134) has more advantages due to lower off-target effects and rapid gene editing (135,136) when compared with plasmid DNA transfection. However, the in vivo delivery formats of mRNA and protein pose certain technical challenges, including inability to globally deliver to the target cells or organs or trigger RNA-sensing innate immune responses in human and murine cells, leading to cytotoxicity (137). Further improvement in the performance of the CRISPR components and means of delivery is necessary to increase the efficiency of in vivo genome editing, such as delivering nanoparticles carrying sgRNA and Cas9 protein (138).

In conclusion, there are some difficulties that need to be overcome for effective use of CRISPR-based gene editing, especially in clinical therapy. Nevertheless, genome editing has become a powerful tool for modifying cell lines and organisms to investigate the biology and pathophysiological mechanisms of various genetic diseases. Further studies will continue in order to improve gene targeting accuracy and efficiency, delivery potentials into particular cells, tissues or organs, detection and modulation efficiencies of activity time and activity area of Cas9 in vivo, and prediction and treatment efficiencies of undesired mutations caused by gene editing. Another important issue is represented by the ethical concern related to the use of CRISPR technology in humans, and the appropriate ethical and regulatory guidelines that must be developed to judge the reasonable use of these tools. There is anticipation that we can integrally understand the intricacy and diversity of CRISPR/Cas in gene editing and thereby improve this technology.

Conflict of Interest statement. None declared.

Funding

This work was supported by the grants from Chinese National Nature Science Foundation (81570932, U1738102) and National Basic Research Program of China (2014CB541705).

References

1

Komor
A.C.
,
Badran
A.H.
,
Liu
D.R.
(
2017
)
CRISPR-based technologies for the manipulation of eukaryotic genomes
.
Cell
,
168
,
20
36
.

2

Makarova
K.S.
,
Zhang
F.
,
Koonin
E.V.
(
2017
)
SnapShot: class 1 CRISPR-Cas systems
.
Cell
,
168
,
946
946.e1
.

3

Makarova
K.S.
,
Zhang
F.
,
Koonin
E.V.
(
2017
)
SnapShot: class 2 CRISPR-Cas Systems
.
Cell
,
168
,
328
328.e1
.

4

Jiang
F.
,
Doudna
J.A.
(
2017
)
CRISPR-Cas9 structures and mechanisms
.
Annu. Rev. Biophys
.,
46
,
505
529
.

5

Miyaoka
Y.
,
Berman
J.R.
,
Cooper
S.B.
,
Mayerl
S.J.
,
Chan
A.H.
,
Zhang
B.
,
Karlin-Neumann
G.A.
,
Conklin
B.R.
(
2016
)
Systematic quantification of HDR and NHEJ reveals effects of locus, nuclease, and cell type on genome-editing
.
Sci. Rep
.,
6
,
23549
.

6

Landrum
M.J.
,
Lee
J.M.
,
Benson
M.
,
Brown
G.
,
Chao
C.
,
Chitipiralla
S.
,
Gu
B.
,
Hart
J.
,
Hoffman
D.
,
Hoover
J.
et al. (
2016
)
ClinVar: public archive of interpretations of clinically relevant variants
.
Nucleic Acids Res
.,
44
,
D862
D868
.

7

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. (
2013
)
Multiplex genome engineering using CRISPR/Cas systems
.
Science
,
339
,
819
823
.

8

Ran
F.A.
,
Hsu
P.D.
,
Wright
J.
,
Agarwala
V.
,
Scott
D.A.
,
Zhang
F.
(
2013
)
Genome engineering using the CRISPR-Cas9 system
.
Nat. Protocol
.,
8
,
2281
2308
.

9

Komor
A.C.
,
Kim
Y.B.
,
Packer
M.S.
,
Zuris
J.A.
,
Liu
D.R.
(
2016
)
Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage
.
Nature
,
533
,
420
424
.

10

Kim
D.
,
Lim
K.
,
Kim
S.T.
,
Yoon
S.H.
,
Kim
K.
,
Ryu
S.M.
,
Kim
J.S.
(
2017
)
Genome-wide target specificities of CRISPR RNA-guided programmable deaminases
.
Nat. Biotechnol
.,
35
,
475
480
.

11

Rees
H.A.
,
Komor
A.C.
,
Yeh
W.H.
,
Caetano-Lopes
J.
,
Warman
M.
,
Edge
A.S.B.
,
Liu
D.R.
(
2017
)
Improving the DNA specificity and applicability of base editing through protein engineering and protein delivery
.
Nat. Commun
.,
8
,
15790
.

12

Komor
A.C.
,
Zhao
K.T.
,
Packer
M.S.
,
Gaudelli
N.M.
,
Waterbury
A.L.
,
Koblan
L.W.
,
Kim
Y.B.
,
Badran
A.H.
,
Liu
D.R.
(
2017
)
Improved base excision repair inhibition and bacteriophage Mu Gam protein yields C: G-to-T: a base editors with higher efficiency and product purity
.
Sci. Adv
.,
3
,
eaao4774
.

13

Nishida
K.
,
Arazoe
T.
,
Yachie
N.
,
Banno
S.
,
Kakimoto
M.
,
Tabata
M.
,
Mochizuki
M.
,
Miyabe
A.
,
Araki
M.
,
Hara
K.Y.
et al. (
2016
)
Targeted nucleotide editing using hybrid prokaryotic and vertebrate adaptive immune systems
.
Science
,
353
,
aaf8729
.

14

Hess
G.T.
,
Fresard
L.
,
Han
K.
,
Lee
C.H.
,
Li
A.
,
Cimprich
K.A.
,
Montgomery
S.B.
,
Bassik
M.C.
(
2016
)
Directed evolution using dCas9-targeted somatic hypermutation in mammalian cells
.
Nat. Methods
,
13
,
1036
1042
.

15

Kim
K.
,
Ryu
S.M.
,
Kim
S.T.
,
Baek
G.
,
Kim
D.
,
Lim
K.
,
Chung
E.
,
Kim
S.
,
Kim
J.S.
(
2017
)
Highly efficient RNA-guided base editing in mouse embryos
.
Nat. Biotechnol
.,
35
,
435
437
.

16

Kuscu
C.
,
Parlak
M.
,
Tufan
T.
,
Yang
J.
,
Szlachta
K.
,
Wei
X.
,
Mammadov
R.
,
Adli
M.
(
2017
)
CRISPR-STOP: gene silencing through base-editing-induced nonsense mutations
.
Nat. Methods
,
14
,
710
712
.

17

Lu
Y.
,
Zhu
J.K.
(
2017
)
Precise editing of a target base in the rice genome using a modified CRISPR/Cas9 system
.
Mol. Plant
,
10
,
523
525
.

18

Ma
Y.
,
Zhang
J.
,
Yin
W.
,
Zhang
Z.
,
Song
Y.
,
Chang
X.
(
2016
)
Targeted AID-mediated mutagenesis (TAM) enables efficient genomic diversification in mammalian cells
.
Nat. Methods
,
13
,
1029
1035
.

19

Yang
L.
,
Briggs
A.W.
,
Chew
W.L.
,
Mali
P.
,
Guell
M.
,
Aach
J.
,
Goodman
D.B.
,
Cox
D.
,
Kan
Y.
,
Lesha
E.
et al. (
2016
)
Engineering and optimising deaminase fusions for genome editing
.
Nat. Commun
.,
7
,
13330
.

20

Zong
Y.
,
Wang
Y.
,
Li
C.
,
Zhang
R.
,
Chen
K.
,
Ran
Y.
,
Qiu
J.L.
,
Wang
D.
,
Gao
C.
(
2017
)
Precise base editing in rice, wheat and maize with a Cas9-cytidine deaminase fusion
.
Nat. Biotechnol
.,
35
,
438
440
.

21

Shimatani
Z.
,
Kashojiya
S.
,
Takayama
M.
,
Terada
R.
,
Arazoe
T.
,
Ishii
H.
,
Teramura
H.
,
Yamamoto
T.
,
Komatsu
H.
,
Miura
K.
et al. (
2017
)
Targeted base editing in rice and tomato using a CRISPR-Cas9 cytidine deaminase fusion
.
Nat. Biotechnol
.,
35
,
441
443
.

22

Billon
P.
,
Bryant
E.E.
,
Joseph
S.A.
,
Nambiar
T.S.
,
Hayward
S.B.
,
Rothstein
R.
,
Ciccia
A.
(
2017
)
CRISPR-mediated base editing enables efficient disruption of eukaryotic genes through induction of STOP codons
.
Mol. Cell
,
67
,
1068
1079.e4
.

23

Kim
J.S.
(
2018
)
Precision genome engineering through adenine and cytosine base editing
.
Nat. Plants
,
4
,
148
151
.

24

Gilbert
L.A.
,
Horlbeck
M.A.
,
Adamson
B.
,
Villalta
J.E.
,
Chen
Y.
,
Whitehead
E.H.
,
Guimaraes
C.
,
Panning
B.
,
Ploegh
H.L.
,
Bassik
M.C.
et al. (
2014
)
Genome-scale CRISPR-mediated control of gene repression and activation
.
Cell
,
159
,
647
661
.

25

Maeder
M.L.
,
Linder
S.J.
,
Cascio
V.M.
,
Fu
Y.
,
Ho
Q.H.
,
Joung
J.K.
(
2013
)
CRISPR RNA-guided activation of endogenous human genes
.
Nat. Methods
,
10
,
977
979
.

26

Perez-Pinera
P.
,
Kocak
D.D.
,
Vockley
C.M.
,
Adler
A.F.
,
Kabadi
A.M.
,
Polstein
L.R.
,
Thakore
P.I.
,
Glass
K.A.
,
Ousterout
D.G.
,
Leong
K.W.
et al. (
2013
)
RNA-guided gene activation by CRISPR-Cas9-based transcription factors
.
Nat. Methods
,
10
,
973
976
.

27

Gilbert
L.A.
,
Larson
M.H.
,
Morsut
L.
,
Liu
Z.
,
Brar
G.A.
,
Torres
S.E.
,
Stern-Ginossar
N.
,
Brandman
O.
,
Whitehead
E.H.
,
Doudna
J.A.
et al. (
2013
)
CRISPR-mediated modular RNA-guided regulation of transcription in eukaryotes
.
Cell
,
154
,
442
451
.

28

Bikard
D.
,
Jiang
W.
,
Samai
P.
,
Hochschild
A.
,
Zhang
F.
,
Marraffini
L.A.
(
2013
)
Programmable repression and activation of bacterial gene expression using an engineered CRISPR-Cas system
.
Nucleic Acids Res
.,
41
,
7429
7437
.

29

Cheng
A.W.
,
Wang
H.
,
Yang
H.
,
Shi
L.
,
Katz
Y.
,
Theunissen
T.W.
,
Rangarajan
S.
,
Shivalila
C.S.
,
Dadon
D.B.
,
Jaenisch
R.
(
2013
)
Multiplexed activation of endogenous genes by CRISPR-on, an RNA-guided transcriptional activator system
.
Cell Res
.,
23
,
1163
1171
.

30

Perez-Pinera
P.
,
Ousterout
D.G.
,
Brunger
J.M.
,
Farin
A.M.
,
Glass
K.A.
,
Guilak
F.
,
Crawford
G.E.
,
Hartemink
A.J.
,
Gersbach
C.A.
(
2013
)
Synergistic and tunable human gene activation by combinations of synthetic transcription factors
.
Nat. Methods
,
10
,
239
242
.

31

Gao
X.
,
Tsang
J.C.
,
Gaba
F.
,
Wu
D.
,
Lu
L.
,
Liu
P.
(
2014
)
Comparison of TALE designer transcription factors and the CRISPR/dCas9 in regulation of gene expression by targeting enhancers
.
Nucleic Acids Res
.,
42
,
e155
.

32

Chavez
A.
,
Scheiman
J.
,
Vora
S.
,
Pruitt
B.W.
,
Tuttle
M.
,
P R Iyer
E.
,
Lin
S.
,
Kiani
S.
,
Guzman
C.D.
,
Wiegand
D.J.
et al. (
2015
)
Highly efficient Cas9-mediated transcriptional programming
.
Nat. Methods
,
12
,
326
328
.

33

Mali
P.
,
Aach
J.
,
Stranges
P.B.
,
Esvelt
K.M.
,
Moosburner
M.
,
Kosuri
S.
,
Yang
L.
,
Church
G.M.
(
2013
)
CAS9 transcriptional activators for target specificity screening and paired nickases for cooperative genome engineering
.
Nat. Biotechnol
.,
31
,
833
838
.

34

Konermann
S.
,
Brigham
M.D.
,
Trevino
A.E.
,
Joung
J.
,
Abudayyeh
O.O.
,
Barcena
C.
,
Hsu
P.D.
,
Habib
N.
,
Gootenberg
J.S.
,
Nishimasu
H.
et al. (
2014
)
Genome-scale transcriptional activation by an engineered CRISPR-Cas9 complex
.
Nature
,
517
,
583
588
.

35

Tanenbaum
M.E.
,
Gilbert
L.A.
,
Qi
L.S.
,
Weissman
J.S.
,
Vale
R.D.
(
2014
)
A protein-tagging system for signal amplification in gene expression and fluorescence imaging
.
Cell
,
159
,
635
646
.

36

Waryah
C.B.
,
Moses
C.
,
Arooj
M.
,
Blancafort
P.
(
2018
)
Zinc Fingers, TALEs, and CRISPR systems: a comparison of tools for epigenome editing
.
Methods Mol. Biol
.,
1767
,
19
63
.

37

Morita
S.
,
Noguchi
H.
,
Horii
T.
,
Nakabayashi
K.
,
Kimura
M.
,
Okamura
K.
,
Sakai
A.
,
Nakashima
H.
,
Hata
K.
,
Nakashima
K.
et al. (
2016
)
Targeted DNA demethylation in vivo using dCas9-peptide repeat and scFv-TET1 catalytic domain fusions
.
Nat. Biotechnol
.,
34
,
1060
1065
.

38

Choudhury
S.R.
,
Cui
Y.
,
Lubecka
K.
,
Stefanska
B.
,
Irudayaraj
J.
(
2016
)
CRISPR-dCas9 mediated TET1 targeting for selective DNA demethylation at BRCA1 promoter
.
Oncotarget
,
7
,
46545
46556
.

39

Liu
X.S.
,
Wu
H.
,
Ji
X.
,
Stelzer
Y.
,
Wu
X.
,
Czauderna
S.
,
Shu
J.
,
Dadon
D.
,
Young
R.A.
,
Jaenisch
R.
(
2016
)
Editing DNA methylation in the mammalian genome
.
Cell
,
167
,
233
247.e17
.

40

Thakore
P.I.
,
Black
J.B.
,
Hilton
I.B.
,
Gersbach
C.A.
(
2016
)
Editing the epigenome: technologies for programmable transcription and epigenetic modulation
.
Nat. Methods
,
13
,
127
137
.

41

Liao
H.-K.
,
Hatanaka
F.
,
Araoka
T.
,
Reddy
P.
,
Wu
M.-Z.
,
Sui
Y.
,
Yamauchi
T.
,
Sakurai
M.
,
O’Keefe
D.D.
,
Núñez-Delicado
E.
et al. (
2017
)
In vivo target gene activation via CRISPR/Cas9-mediated trans-epigenetic modulation
.
Cell
,
171
,
1495
1507.e15
.

42

Chen
B.
,
Gilbert
L.A.
,
Cimini
B.A.
,
Schnitzbauer
J.
,
Zhang
W.
,
Li
G.W.
,
Park
J.
,
Blackburn
E.H.
,
Weissman
J.S.
,
Qi
L.S.
et al. (
2013
)
Dynamic imaging of genomic loci in living human cells by an optimized CRISPR/Cas system
.
Cell
,
155
,
1479
1491
.

43

Kearns
N.A.
,
Pham
H.
,
Tabak
B.
,
Genga
R.M.
,
Silverstein
N.J.
,
Garber
M.
,
Maehr
R.
(
2015
)
Functional annotation of native enhancers with a Cas9-histone demethylase fusion
.
Nat. Methods
,
12
,
401
403
.

44

Kwon
D.Y.
,
Zhao
Y.T.
,
Lamonica
J.M.
,
Zhou
Z.
(
2017
)
Locus-specific histone deacetylation using a synthetic CRISPR-Cas9-based HDAC
.
Nat. Commun
.,
8
,
15315
.

45

Lei
Y.
,
Zhang
X.
,
Su
J.
,
Jeong
M.
,
Gundry
M.C.
,
Huang
Y.H.
,
Zhou
Y.
,
Li
W.
,
Goodell
M.A.
(
2017
)
Targeted DNA methylation in vivo using an engineered dCas9-MQ1 fusion protein
.
Nat. Commun
.,
8
,
16026
.

46

Huang
Y.H.
,
Su
J.
,
Lei
Y.
,
Brunetti
L.
,
Gundry
M.C.
,
Zhang
X.
,
Jeong
M.
,
Li
W.
,
Goodell
M.A.
(
2017
)
DNA epigenome editing using CRISPR-Cas SunTag-directed DNMT3A
.
Genome Biol
.,
18
,
176
.

47

Konermann
S.
,
Brigham
M.D.
,
Trevino
A.
,
Hsu
P.D.
,
Heidenreich
M.
,
Cong
L.
,
Platt
R.J.
,
Scott
D.A.
,
Church
G.M.
,
Zhang
F.
(
2013
)
Optical control of mammalian endogenous transcription and epigenetic states
.
Nature
,
500
,
472
476
.

48

Klann
T.S.
,
Black
J.B.
,
Chellappan
M.
,
Safi
A.
,
Song
L.
,
Hilton
I.B.
,
Crawford
G.E.
,
Reddy
T.E.
,
Gersbach
C.A.
(
2017
)
CRISPR-Cas9 epigenome editing enables high-throughput screening for functional regulatory elements in the human genome
.
Nat. Biotechnol
.,
35
,
561
568
.

49

Paddison
P.J.
,
Silva
J.M.
,
Conklin
D.S.
,
Schlabach
M.
,
Li
M.
,
Aruleba
S.
,
Balija
V.
,
O'Shaughnessy
A.
,
Gnoj
L.
,
Scobie
K.
et al. (
2004
)
A resource for large-scale RNA-interference-based screens in mammals
.
Nature
,
428
,
427
431
.

50

Root
D.E.
,
Hacohen
N.
,
Hahn
W.C.
,
Lander
E.S.
,
Sabatini
D.M.
(
2006
)
Genome-scale loss-of-function screening with a lentiviral RNAi library
.
Nat. Methods
,
3
,
715
719
.

51

Silva
J.M.
,
Li
M.Z.
,
Chang
K.
,
Ge
W.
,
Golding
M.C.
,
Rickles
R.J.
,
Siolas
D.
,
Hu
G.
,
Paddison
P.J.
,
Schlabach
M.R.
et al. (
2005
)
Second-generation shRNA libraries covering the mouse and human genomes
.
Nat. Genet
.,
37
,
1281
1288
.

52

Jackson
A.L.
,
Bartz
S.R.
,
Schelter
J.
,
Kobayashi
S.V.
,
Burchard
J.
,
Mao
M.
,
Li
B.
,
Cavet
G.
,
Linsley
P.S.
(
2003
)
Expression profiling reveals off-target gene regulation by RNAi
.
Nat. Biotechnol
.,
21
,
635
637
.

53

Jackson
A.L.
,
Linsley
P.S.
(
2010
)
Recognizing and avoiding siRNA off-target effects for target identification and therapeutic application
.
Nat. Rev. Drug Discov
.,
9
,
57
67
.

54

Schumann
K.
,
Lin
S.
,
Boyer
E.
,
Simeonov
D.R.
,
Subramaniam
M.
,
Gate
R.E.
,
Haliburton
G.E.
,
Ye
C.J.
,
Bluestone
J.A.
,
Doudna
J.A.
et al. (
2015
)
Generation of knock-in primary human T cells using Cas9 ribonucleoproteins
.
Proc. Natl. Acad. Sci. U. S. A
.,
112
,
10437
10442
.

55

Korkmaz
G.
,
Lopes
R.
,
Ugalde
A.P.
,
Nevedomskaya
E.
,
Han
R.
,
Myacheva
K.
,
Zwart
W.
,
Elkon
R.
,
Agami
R.
(
2016
)
Functional genetic screens for enhancer elements in the human genome using CRISPR-Cas9
.
Nat. Biotechnol
.,
34
,
192
198
.

56

Cox
D.B.T.
,
Gootenberg
J.S.
,
Abudayyeh
O.O.
,
Franklin
B.
,
Kellner
M.J.
,
Joung
J.
,
Zhang
F.
(
2017
)
RNA editing with CRISPR-Cas13
.
Science
,
358
,
1019
1027
.

57

Gaudelli
N.M.
,
Komor
A.C.
,
Rees
H.A.
,
Packer
M.S.
,
Badran
A.H.
,
Bryson
D.I.
,
Liu
D.R.
(
2017
)
Programmable base editing of A*T to G*C in genomic DNA without DNA cleavage
.
Nature
,
551
,
464
471
.

58

DeJesus
R.
,
Moretti
F.
,
McAllister
G.
,
Wang
Z.
,
Bergman
P.
,
Liu
S.
,
Frias
E.
,
Alford
J.
,
Reece-Hoyes
J.S.
,
Lindeman
A.
et al. (
2016
)
Functional CRISPR screening identifies the ufmylation pathway as a regulator of SQSTM1/p62
.
Elife
,
5

59

Adamson
B.
,
Norman
T.M.
,
Jost
M.
,
Cho
M.Y.
,
Nunez
J.K.
,
Chen
Y.
,
Villalta
J.E.
,
Gilbert
L.A.
,
Horlbeck
M.A.
,
Hein
M.Y.
et al. (
2016
)
A multiplexed single-cell CRISPR screening platform enables systematic dissection of the unfolded protein response
.
Cell
,
167
,
1867
1882.e21
.

60

Dixit
A.
,
Parnas
O.
,
Li
B.
,
Chen
J.
,
Fulco
C.P.
,
Jerby-Arnon
L.
,
Marjanovic
N.D.
,
Dionne
D.
,
Burks
T.
,
Raychowdhury
R.
et al. (
2016
)
Perturb-Seq: dissecting molecular circuits with scalable single-cell RNA profiling of pooled genetic screens
.
Cell
,
167
,
1853
1866.e17
.

61

Jaitin
D.A.
,
Weiner
A.
,
Yofe
I.
,
Lara-Astiaso
D.
,
Keren-Shaul
H.
,
David
E.
,
Salame
T.M.
,
Tanay
A.
,
van Oudenaarden
A.
,
Amit
I.
(
2016
)
Dissecting immune circuits by linking CRISPR-pooled screens with single-cell RNA-Seq
.
Cell
,
167
,
1883
1896.e15
.

62

Steinhart
Z.
,
Pavlovic
Z.
,
Chandrashekhar
M.
,
Hart
T.
,
Wang
X.
,
Zhang
X.
,
Robitaille
M.
,
Brown
K.R.
,
Jaksani
S.
,
Overmeer
R.
et al. (
2017
)
Genome-wide CRISPR screens reveal a Wnt-FZD5 signaling circuit as a druggable vulnerability of RNF43-mutant pancreatic tumors
.
Nat. Med
.,
23
,
60
68
.

63

Han
K.
,
Jeng
E.E.
,
Hess
G.T.
,
Morgens
D.W.
,
Li
A.
,
Bassik
M.C.
(
2017
)
Synergistic drug combinations for cancer identified in a CRISPR screen for pairwise genetic interactions
.
Nat. Biotechnol
.,
35
,
463
474
.

64

Park
R.J.
,
Wang
T.
,
Koundakjian
D.
,
Hultquist
J.F.
,
Lamothe-Molina
P.
,
Monel
B.
,
Schumann
K.
,
Yu
H.
,
Krupzcak
K.M.
,
Garcia-Beltran
W.
et al. (
2017
)
A genome-wide CRISPR screen identifies a restricted set of HIV host dependency factors
.
Nat. Genet
.,
49
,
193
203
.

65

Manguso
R.T.
,
Pope
H.W.
,
Zimmer
M.D.
,
Brown
F.D.
,
Yates
K.B.
,
Miller
B.C.
,
Collins
N.B.
,
Bi
K.
,
LaFleur
M.W.
,
Juneja
V.R.
et al. (
2017
)
In vivo CRISPR screening identifies Ptpn2 as a cancer immunotherapy target
.
Nature
,
547
,
413
418
.

66

Song
C.Q.
,
Li
Y.
,
Mou
H.
,
Moore
J.
,
Park
A.
,
Pomyen
Y.
,
Hough
S.
,
Kennedy
Z.
,
Fischer
A.
,
Yin
H.
et al. (
2017
)
Genome-wide CRISPR screen identifies regulators of mitogen-activated protein kinase as suppressors of liver tumors in mice
.
Gastroenterology
,
152
,
1161
1173.e1
.

67

Yamauchi
T.
,
Masuda
T.
,
Canver
M.C.
,
Seiler
M.
,
Semba
Y.
,
Shboul
M.
,
Al-Raqad
M.
,
Maeda
M.
,
Schoonenberg
V.A.C.
,
Cole
M.A.
et al. (
2018
)
Genome-wide CRISPR-Cas9 screen identifies leukemia-specific dependence on a pre-mRNA metabolic pathway regulated by DCPS
.
Cancer Cell
,
33
,
386
400.e5
.

68

Brunen
D.
,
de Vries
R.C.
,
Lieftink
C.
,
Beijersbergen
R.L.
,
Bernards
R.
(
2018
) PIM kinases are a potential prognostic biomarker and therapeutic target in neuroblastoma. Mol. Cancer Ther.,
17
,
849
857
.

69

Zhang
S.
,
Zhang
M.
,
Jing
Y.
,
Yin
X.
,
Ma
P.
,
Zhang
Z.
,
Wang
X.
,
Di
W.
,
Zhuang
G.
(
2018
)
Deubiquitinase USP13 dictates MCL1 stability and sensitivity to BH3 mimetic inhibitors
.
Nat. Commun
.,
9
,
215
.

70

Kurata
M.
,
Rathe
S.K.
,
Bailey
N.J.
,
Aumann
N.K.
,
Jones
J.M.
,
Veldhuijzen
G.W.
,
Moriarity
B.S.
,
Largaespada
D.A.
(
2016
)
Using genome-wide CRISPR library screening with library resistant DCK to find new sources of Ara-C drug resistance in AML
.
Sci. Rep
.,
6
,
36199
.

71

Sanjana
N.E.
,
Wright
J.
,
Zheng
K.
,
Shalem
O.
,
Fontanillas
P.
,
Joung
J.
,
Cheng
C.
,
Regev
A.
,
Zhang
F.
(
2016
)
High-resolution interrogation of functional elements in the noncoding genome
.
Science
,
353
,
1545
1549
.

72

Diao
Y.
,
Li
B.
,
Meng
Z.
,
Jung
I.
,
Lee
A.Y.
,
Dixon
J.
,
Maliskova
L.
,
Guan
K.L.
,
Shen
Y.
,
Ren
B.
(
2016
)
A new class of temporarily phenotypic enhancers identified by CRISPR/Cas9-mediated genetic screening
.
Genome Res
.,
26
,
397
405
.

73

Qiu
X.Y.
,
Zhu
L.Y.
,
Zhu
C.S.
,
Ma
J.X.
,
Hou
T.
,
Wu
X.M.
,
Xie
S.S.
,
Min
L.
,
Tan
D.A.
,
Zhang
D.Y.
et al. (
2018
)
Highly effective and low-cost microRNA detection with CRISPR-Cas9
.
ACS Synth. Biol
.,
7
,
807
813
.

74

Chow
R.D.
,
Guzman
C.D.
,
Wang
G.
,
Schmidt
F.
,
Youngblood
M.W.
,
Ye
L.
,
Errami
Y.
,
Dong
M.B.
,
Martinez
M.A.
,
Zhang
S.
et al. (
2017
)
AAV-mediated direct in vivo CRISPR screen identifies functional suppressors in glioblastoma
.
Nat. Neurosci
.,
20
,
1329
1341
.

75

Wang
G.
,
Chow
R.D.
,
Ye
L.
,
Guzman
C.D.
,
Dai
X.
,
Dong
M.B.
,
Zhang
F.
,
Sharp
P.A.
,
Platt
R.J.
,
Chen
S.
(
2018
)
Mapping a functional cancer genome atlas of tumor suppressors in mouse liver using AAV-CRISPR-mediated direct in vivo screening
.
Sci. Adv
.,
4
,
eaao5508
.

76

Ren
J.
,
Liu
X.
,
Fang
C.
,
Jiang
S.
,
June
C.H.
,
Zhao
Y.
(
2017
)
Multiplex genome editing to generate universal CAR T cells resistant to PD1 inhibition
.
Clin. Cancer Res
.,
23
,
2255
2266
.

77

Maude
S.L.
,
Frey
N.
,
Shaw
P.A.
,
Aplenc
R.
,
Barrett
D.M.
,
Bunin
N.J.
,
Chew
A.
,
Gonzalez
V.E.
,
Zheng
Z.
,
Lacey
S.F.
et al. (
2014
)
Chimeric antigen receptor T cells for sustained remissions in leukemia
.
N. Engl. J. Med
.,
371
,
1507
1517
.

78

Safari
F.
,
Farajnia
S.
,
Arya
M.
,
Zarredar
H.
,
Nasrolahi
A.
(
2018
)
CRISPR and personalized Treg therapy: new insights into the treatment of rheumatoid arthritis
.
Immunopharmacol. Immunotoxicol
.,
1
11
.

79

Rupp
L.J.
,
Schumann
K.
,
Roybal
K.T.
,
Gate
R.E.
,
Ye
C.J.
,
Lim
W.A.
,
Marson
A.
(
2017
)
CRISPR/Cas9-mediated PD-1 disruption enhances anti-tumor efficacy of human chimeric antigen receptor T cells
.
Sci. Rep
.,
7
,
737
.

80

Gomes-Silva
D.
,
Srinivasan
M.
,
Sharma
S.
,
Lee
C.M.
,
Wagner
D.L.
,
Davis
T.H.
,
Rouce
R.H.
,
Bao
G.
,
Brenner
M.K.
,
Mamonkin
M.
(
2017
)
CD7-edited T cells expressing a CD7-specific CAR for the therapy of T-cell malignancies
.
Blood
,
130
,
285
296
.

81

Hendel
A.
,
Bak
R.O.
,
Clark
J.T.
,
Kennedy
A.B.
,
Ryan
D.E.
,
Roy
S.
,
Steinfeld
I.
,
Lunstad
B.D.
,
Kaiser
R.J.
,
Wilkens
A.B.
et al. (
2015
)
Chemically modified guide RNAs enhance CRISPR-Cas genome editing in human primary cells
.
Nat. Biotechnol
.,
33
,
985
989
.

82

Porter
D.L.
,
Levine
B.L.
,
Kalos
M.
,
Bagg
A.
,
June
C.H.
(
2011
)
Chimeric antigen receptor-modified T cells in chronic lymphoid leukemia
.
N. Engl. J. Med
.,
365
,
725
733
.

83

Yi
L.
,
Li
J.
(
2016
)
CRISPR-Cas9 therapeutics in cancer: promising strategies and present challenges
.
Biochim. Biophys. Acta
,
1866
,
197
207
.

84

Liu
X.
,
Zhang
Y.
,
Cheng
C.
,
Cheng
A.W.
,
Zhang
X.
,
Li
N.
,
Xia
C.
,
Wei
X.
,
Liu
X.
,
Wang
H.
(
2017
)
CRISPR-Cas9-mediated multiplex gene editing in CAR-T cells
.
Cell Res
.,
27
,
154
157
.

85

Ren
J.
,
Zhang
X.
,
Liu
X.
,
Fang
C.
,
Jiang
S.
,
June
C.H.
,
Zhao
Y.
(
2017
)
A versatile system for rapid multiplex genome-edited CAR T cell generation
.
Oncotarget
,
8
,
17002
17011
.

86

Eyquem
J.
,
Mansilla-Soto
J.
,
Giavridis
T.
,
van der Stegen
S.J.
,
Hamieh
M.
,
Cunanan
K.M.
,
Odak
A.
,
Gonen
M.
,
Sadelain
M.
(
2017
)
Targeting a CAR to the TRAC locus with CRISPR/Cas9 enhances tumour rejection
.
Nature
,
543
,
113
117
.

87

Barrangou
R.
,
Doudna
J.A.
(
2016
)
Applications of CRISPR technologies in research and beyond
.
Nat. Biotechnol
.,
34
,
933
941
.

88

Allende
M.L.
,
Cook
E.K.
,
Larman
B.C.
,
Nugent
A.
,
Brady
J.M.
,
Golebiowski
D.
,
Sena-Esteves
M.
,
Tifft
C.J.
,
Proia
R.L.
(
2018
)
Cerebral organoids derived from Sandhoff disease-induced pluripotent stem cells exhibit impaired neurodifferentiation
.
J. Lipid Res
.,
59
,
550
563
.

89

Yamamoto
Y.
,
Makiyama
T.
,
Harita
T.
,
Sasaki
K.
,
Wuriyanghai
Y.
,
Hayano
M.
,
Nishiuchi
S.
,
Kohjitani
H.
,
Hirose
S.
,
Chen
J.
et al. (
2017
)
Allele-specific ablation rescues electrophysiological abnormalities in a human iPS cell model of long-QT syndrome with a CALM2 mutation
.
Hum. Mol. Genet
.,
26
,
1670
1677
.

90

Saito
A.
,
Ooki
A.
,
Nakamura
T.
,
Onodera
S.
,
Hayashi
K.
,
Hasegawa
D.
,
Okudaira
T.
,
Watanabe
K.
,
Kato
H.
,
Onda
T.
et al. (
2018
)
Targeted reversion of induced pluripotent stem cells from patients with human cleidocranial dysplasia improves bone regeneration in a rat calvarial bone defect model
.
Stem Cell Res. Ther
.,
9
,
12
.

91

Kim
S.
,
Yun
S.P.
,
Lee
S.
,
Umanah
G.E.
,
Bandaru
V.V.R.
,
Yin
X.
,
Rhee
P.
,
Karuppagounder
S.S.
,
Kwon
S.H.
,
Lee
H.
et al. (
2018
)
GBA1 deficiency negatively affects physiological alpha-synuclein tetramers and related multimers
.
Proc. Natl. Acad. Sci. U. S. A
.,
115
,
798
803
.

92

Romero-Moya
D.
,
Santos-Ocana
C.
,
Castano
J.
,
Garrabou
G.
,
Rodriguez-Gomez
J.A.
,
Ruiz-Bonilla
V.
,
Bueno
C.
,
Gonzalez-Rodriguez
P.
,
Giorgetti
A.
,
Perdiguero
E.
et al. (
2017
)
Genetic rescue of mitochondrial and skeletal muscle impairment in an induced pluripotent. Stem cells model of coenzyme Q10 deficiency
.
Stem Cell
,
35
,
1687
1703
.

93

Hashem
S.I.
,
Murphy
A.N.
,
Divakaruni
A.S.
,
Klos
M.L.
,
Nelson
B.C.
,
Gault
E.C.
,
Rowland
T.J.
,
Perry
C.N.
,
Gu
Y.
,
Dalton
N.D.
et al. (
2017
)
Impaired mitophagy facilitates mitochondrial damage in Danon disease
.
J. Mol. Cell Cardiol
.,
108
,
86
94
.

94

Hu
L.
,
Du
L.
,
Zhao
Y.
,
Li
W.
,
Ouyang
Q.
,
Zhou
D.
,
Lu
G.
,
Lin
G.
(
2017
)
Modeling Glanzmann thrombasthenia using patient specific iPSCs and restoring platelet aggregation function by CD41 overexpression
.
Stem Cell Res
.,
20
,
14
20
.

95

Meneghini
V.
,
Frati
G.
,
Sala
D.
,
De Cicco
S.
,
Luciani
M.
,
Cavazzin
C.
,
Paulis
M.
,
Mentzen
W.
,
Morena
F.
,
Giannelli
S.
et al. (
2017
)
Generation of human induced pluripotent stem cell-derived bona fide neural stem cells for ex vivo gene therapy of metachromatic leukodystrophy
.
Stem Cells Transl. Med
.,
6
,
352
368
.

96

Uusi-Rauva
K.
,
Blom
T.
,
von Schantz-Fant
C.
,
Blom
T.
,
Jalanko
A.
,
Kyttala
A.
(
2017
)
Induced pluripotent stem cells derived from a CLN5 patient manifest phenotypic characteristics of neuronal ceroid lipofuscinoses
.
Int. J. Mol. Sci
.,
18
,
955
.

97

Rabenstein
M.
,
Peter
F.
,
Joost
S.
,
Trilck
M.
,
Rolfs
A.
,
Frech
M.J.
(
2017
)
Decreased calcium flux in Niemann-Pick type C1 patient-specific iPSC-derived neurons due to higher amount of calcium-impermeable AMPA receptors
.
Mol. Cell Neurosci
.,
83
,
27
36
.

98

Nevin
Z.S.
,
Factor
D.C.
,
Karl
R.T.
,
Douvaras
P.
,
Laukka
J.
,
Windrem
M.S.
,
Goldman
S.A.
,
Fossati
V.
,
Hobson
G.M.
,
Tesar
P.J.
(
2017
)
Modeling the mutational and phenotypic landscapes of Pelizaeus-Merzbacher disease with human iPSC-derived oligodendrocytes
.
Am. J. Hum. Genet
.,
100
,
617
634
.

99

Sato
Y.
,
Kobayashi
H.
,
Higuchi
T.
,
Shimada
Y.
,
Ida
H.
,
Ohashi
T.
(
2017
)
Metabolomic profiling of pompe disease-induced pluripotent stem cell-derived cardiomyocytes reveals that oxidative stress is associated with cardiac and skeletal muscle pathology
.
Stem Cells Transl. Med
.,
6
,
31
39
.

100

Burnett
L.C.
,
LeDuc
C.A.
,
Sulsona
C.R.
,
Paull
D.
,
Eddiry
S.
,
Levy
B.
,
Salles
J.P.
,
Tauber
M.
,
Driscoll
D.J.
,
Egli
D.
,
Leibel
R.L.
(
2016
)
Induced pluripotent stem cells (iPSC) created from skin fibroblasts of patients with Prader-Willi syndrome (PWS) retain the molecular signature of PWS
.
Stem Cell Res
.,
17
,
526
530
.

101

Chailangkarn
T.
,
Muotri
A.R.
(
2017
)
Modeling Williams syndrome with induced pluripotent stem cells
.
Neurogenesis (Austin)
,
4
,
e1283187
.

102

Ramsden
C.M.
,
Nommiste
B.
,
R. Lane
A.
,
Carr
A.-J.F.
,
Powner
M.B.
,
J. K. Smart
M.
,
Chen
L.L.
,
Muthiah
M.N.
,
Webster
A.R.
,
Moore
A.T.
et al. (
2017
)
Rescue of the MERTK phagocytic defect in a human iPSC disease model using translational read-through inducing drugs
.
Sci. Rep
.,
7
,
51
.

103

Francis
K.R.
,
Ton
A.N.
,
Xin
Y.
,
O'Halloran
P.E.
,
Wassif
C.A.
,
Malik
N.
,
Williams
I.M.
,
Cluzeau
C.V.
,
Trivedi
N.S.
,
Pavan
W.J.
et al. (
2016
)
Modeling Smith-Lemli-Opitz syndrome with induced pluripotent stem cells reveals a causal role for Wnt/beta-catenin defects in neuronal cholesterol synthesis phenotypes
.
Nat. Med
.,
22
,
388
396
.

104

Aguisanda
F.
,
Yeh
C.D.
,
Chen
C.Z.
,
Li
R.
,
Beers
J.
,
Zou
J.
,
Thorne
N.
,
Zheng
W.
(
2017
)
Neural stem cells for disease modeling of Wolman disease and evaluation of therapeutics
.
Orphanet. J. Rare Dis
.,
12
,
120
.

105

Pittermann
E.
,
Lachmann
N.
,
MacLean
G.
,
Emmrich
S.
,
Ackermann
M.
,
Gohring
G.
,
Schlegelberger
B.
,
Welte
K.
,
Schambach
A.
,
Heckl
D.
et al. (
2017
)
Gene correction of HAX1 reversed Kostmann disease phenotype in patient-specific induced pluripotent stem cells
.
Blood Adv
.,
1
,
903
914
.

106

Megaw
R.
,
Abu-Arafeh
H.
,
Jungnickel
M.
,
Mellough
C.
,
Gurniak
C.
,
Witke
W.
,
Zhang
W.
,
Khanna
H.
,
Mill
P.
,
Dhillon
B.
et al. (
2017
)
Gelsolin dysfunction causes photoreceptor loss in induced pluripotent cell and animal retinitis pigmentosa models
.
Nat. Commun
.,
8
,
271
.

107

Parfitt
D.A.
,
Lane
A.
,
Ramsden
C.M.
,
Carr
A.J.
,
Munro
P.M.
,
Jovanovic
K.
,
Schwarz
N.
,
Kanuga
N.
,
Muthiah
M.N.
,
Hull
S.
et al. (
2016
)
Identification and correction of mechanisms underlying inherited blindness in human iPSC-derived optic cups
.
Cell Stem Cell
,
18
,
769
781
.

108

Hadoux
J.
,
Desterke
C.
,
Feraud
O.
,
Guibert
M.
,
De Rose
R.F.
,
Opolon
P.
,
Divers
D.
,
Gobbo
E.
,
Griscelli
F.
,
Schlumberger
M.
et al. (
2018
)
Transcriptional landscape of a RET(C634Y)-mutated iPSC and its CRISPR-corrected isogenic control reveals the putative role of EGR1 transcriptional program in the development of multiple endocrine neoplasia type 2A-associated cancers
.
Stem Cell Res
.,
26
,
8
16
.

109

Li
Y.
,
Jin
C.
,
Bai
H.
,
Gao
Y.
,
Sun
S.
,
Chen
L.
,
Qin
L.
,
Liu
P.P.
,
Cheng
L.
,
Wang
Q.F.
(
2018
)
Human NOTCH4 is a key target of RUNX1 in megakaryocytic differentiation
.
Blood
,
131
,
191
201
.

110

Heman-Ackah
S.M.
,
Manzano
R.
,
Hoozemans
J.J.M.
,
Scheper
W.
,
Flynn
R.
,
Haerty
W.
,
Cowley
S.A.
,
Bassett
A.R.
,
Wood
M.J.A.
(
2017
)
Alpha-synuclein induces the unfolded protein response in Parkinson's disease SNCA triplication iPSC-derived neurons
.
Hum. Mol. Genet
.,
26
,
4441
4450
.

111

Pires
C.
,
Schmid
B.
,
Petræus
C.
,
Poon
A.
,
Nimsanor
N.
,
Nielsen
T.T.
,
Waldemar
G.
,
Hjermind
L.E.
,
Nielsen
J.E.
,
Hyttel
P.
et al. (
2016
)
Generation of a gene-corrected isogenic control cell line from an Alzheimer's disease patient iPSC line carrying a A79V mutation in PSEN1
.
Stem Cell Res
.,
17
,
285
288
.

112

He
Q.
,
Wang
H.H.
,
Cheng
T.
,
Yuan
W.P.
,
Ma
Y.P.
,
Jiang
Y.P.
,
Ren
Z.H.
(
2017
)
Genetic correction and hepatic differentiation of hemophilia B-specific human induced pluripotent stem cells
.
Chin. Med. Sci. J
.,
32
,
135
144
.

113

Tang
Z.H.
,
Chen
J.R.
,
Zheng
J.
,
Shi
H.S.
,
Ding
J.
,
Qian
X.D.
,
Zhang
C.
,
Chen
J.L.
,
Wang
C.C.
,
Li
L.
et al. (
2016
)
Genetic correction of induced pluripotent stem cells from a deaf patient with MYO7A mutation results in morphologic and functional recovery of the derived hair cell-like cells
.
Stem Cells Transl. Med
.,
5
,
561
571
.

114

Chen
J.R.
,
Tang
Z.H.
,
Zheng
J.
,
Shi
H.S.
,
Ding
J.
,
Qian
X.D.
,
Zhang
C.
,
Chen
J.L.
,
Wang
C.C.
,
Li
L.
et al. (
2016
)
Effects of genetic correction on the differentiation of hair cell-like cells from iPSCs with MYO15A mutation
.
Cell Death Differ
.,
23
,
1347
1357
.

115

Guo
D.
,
Liu
H.
,
Gao
G.
,
Liu
Y.
,
Zhuang
Y.
,
Yang
F.
,
Wang
K.
,
Zhou
T.
,
Qin
D.
,
Hong
L.
et al. (
2017
)
Creating a patient carried Men1 gene point mutation on wild type iPSCs locus mediated by CRISPR/Cas9 and ssODN
.
Stem Cell Res
.,
18
,
67
69
.

116

Long
C.
,
Li
H.
,
Tiburcy
M.
,
Rodriguez-Caycedo
C.
,
Kyrychenko
V.
,
Zhou
H.
,
Zhang
Y.
,
Min
Y.L.
,
Shelton
J.M.
,
Mammen
P.P.A.
et al. (
2018
)
Correction of diverse muscular dystrophy mutations in human engineered heart muscle by single-site genome editing
.
Sci. Adv
.,
4
,
eaap9004
.

117

Kraiczy
J.
,
Nayak
K.M.
,
Howell
K.J.
,
Ross
A.
,
Forbester
J.
,
Salvestrini
C.
,
Mustata
R.
,
Perkins
S.
,
Andersson-Rolf
A.
,
Leenen
E.
et al. (
2017
) DNA methylation defines regional identity of human intestinal epithelial organoids and undergoes dynamic changes during development. Gut., doi: 10.1136/gutjnl-2017-314817.

118

Arber
C.
,
Lovejoy
C.
,
Wray
S.
(
2017
)
Stem cell models of Alzheimer's disease: progress and challenges
.
Alzheimers Res. Ther
.,
9
,
42
.

119

Deng
W.L.
,
Gao
M.L.
,
Lei
X.L.
,
Lv
J.N.
,
Zhao
H.
,
He
K.W.
,
Xia
X.X.
,
Li
L.Y.
,
Chen
Y.C.
,
Li
Y.P.
et al. (
2018
)
Gene correction reverses ciliopathy and photoreceptor loss in iPSC-derived retinal organoids from retinitis pigmentosa patients
.
Stem Cell Rep
., doi: 10.1016/j.stemcr.2018.02.003.

120

Mandai
M.
,
Kurimoto
Y.
,
Takahashi
M.
(
2017
)
Autologous induced stem-cell-derived retinal cells for macular degeneration
.
N. Engl. J. Med
.,
377
,
792
793
.

121

Tang
L.
,
Zeng
Y.
,
Du
H.
,
Gong
M.
,
Peng
J.
,
Zhang
B.
,
Lei
M.
,
Zhao
F.
,
Wang
W.
,
Li
X.
et al. (
2017
)
CRISPR/Cas9-mediated gene editing in human zygotes using Cas9 protein
.
Mol. Genet. Genomics
,
292
,
525
533
.

122

Kang
X.
,
He
W.
,
Huang
Y.
,
Yu
Q.
,
Chen
Y.
,
Gao
X.
,
Sun
X.
,
Fan
Y.
(
2016
)
Introducing precise genetic modifications into human 3PN embryos by CRISPR/Cas-mediated genome editing
.
J. Assist. Reprod. Genet
.,
33
,
581
588
.

123

Ma
H.
,
Marti-Gutierrez
N.
,
Park
S.W.
,
Wu
J.
,
Lee
Y.
,
Suzuki
K.
,
Koski
A.
,
Ji
D.
,
Hayama
T.
,
Ahmed
R.
et al. (
2017
)
Correction of a pathogenic gene mutation in human embryos
.
Nature
,
548
,
413
419
.

124

Liang
P.
,
Ding
C.
,
Sun
H.
,
Xie
X.
,
Xu
Y.
,
Zhang
X.
,
Sun
Y.
,
Xiong
Y.
,
Ma
W.
,
Liu
Y.
et al. (
2017
)
Correction of beta-thalassemia mutant by base editor in human embryos
.
Protein Cell
,
8
,
811
822
.

125

Zhou
C.
,
Zhang
M.
,
Wei
Y.
,
Sun
Y.
,
Sun
Y.
,
Pan
H.
,
Yao
N.
,
Zhong
W.
,
Li
Y.
,
Li
W.
et al. (
2017
)
Highly efficient base editing in human tripronuclear zygotes
.
Protein Cell
,
8
,
772
775
.

126

Schenkwein
D.
,
Yla-Herttuala
S.
(
2018
)
Gene editing of human embryos with CRISPR/Cas9: great promise coupled with important caveats
.
Mol. Ther
.,
26
,
659
660
.

127

Cyranoski
D.
,
Reardon
S.
(
2015
)
Embryo editing sparks epic debate
.
Nature
,
520
,
593
594
.

128

National Academies of Sciences, Engineering, and Medicine
. (
2017
)
Human Genome Editing: Science, Ethics, and Governance
The National Academies Press
,
Washington, DC
.

129

Hu
J.H.
,
Miller
S.M.
,
Geurts
M.H.
,
Tang
W.
,
Chen
L.
,
Sun
N.
,
Zeina
C.M.
,
Gao
X.
,
Rees
H.A.
,
Lin
Z.
et al. (
2018
)
Evolved Cas9 variants with broad PAM compatibility and high DNA specificity
.
Nature
,
556
,
57
63
.

130

Kim
I.
,
Jeong
M.
,
Ka
D.
,
Han
M.
,
Kim
N.K.
,
Bae
E.
,
Suh
J.Y.
(
2018
)
Solution structure and dynamics of anti-CRISPR AcrIIA4, the Cas9 inhibitor
.
Sci. Rep
.,
8
,
3883
.

131

Rauch
B.J.
,
Silvis
M.R.
,
Hultquist
J.F.
,
Waters
C.S.
,
McGregor
M.J.
,
Krogan
N.J.
,
Bondy-Denomy
J.
(
2017
)
Inhibition of CRISPR-Cas9 with bacteriophage proteins
.
Cell
,
168
,
150
158.e10
.

132

Sontheimer
E.J.
,
Davidson
A.R.
(
2017
)
Inhibition of CRISPR-Cas systems by mobile genetic elements
.
Curr. Opin. Microbiol
.,
37
,
120
127
.

133

Kay
M.A.
(
2011
)
State-of-the-art gene-based therapies: the road ahead
.
Nat. Rev. Genet
.,
12
,
316
328
.

134

Kouranova
E.
,
Forbes
K.
,
Zhao
G.
,
Warren
J.
,
Bartels
A.
,
Wu
Y.
,
Cui
X.
(
2016
)
CRISPRs for optimal targeting: delivery of CRISPR components as DNA, RNA, and protein into cultured cells and single-cell embryos
.
Hum. Gene Ther
.,
27
,
464
475
.

135

Staahl
B.T.
,
Benekareddy
M.
,
Coulon-Bainier
C.
,
Banfal
A.A.
,
Floor
S.N.
,
Sabo
J.K.
,
Urnes
C.
,
Munares
G.A.
,
Ghosh
A.
,
Doudna
J.A.
(
2017
)
Efficient genome editing in the mouse brain by local delivery of engineered Cas9 ribonucleoprotein complexes
.
Nat. Biotechnol
.,
35
,
431
434
.

136

Kim
S.
,
Kim
D.
,
Cho
S.W.
,
Kim
J.
,
Kim
J.S.
(
2014
)
Highly efficient RNA-guided genome editing in human cells via delivery of purified Cas9 ribonucleoproteins
.
Genome Res
.,
24
,
1012
1019
.

137

Kim
S.
,
Koo
T.
,
Jee
H.G.
,
Cho
H.Y.
,
Lee
G.
,
Lim
D.G.
,
Shin
H.S.
,
Kim
J.S.
(
2018
)
CRISPR RNAs trigger innate immune responses in human cells
.
Genome Res
., doi: 10.1101/gr.231936.117.

138

Gibson
G.J.
,
Yang
M.
(
2017
)
What rheumatologists need to know about CRISPR/Cas9
.
Nat. Rev. Rheumatol
.,
13
,
205
216
.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/about_us/legal/notices)