Next Article in Journal
GenomeBits Characterization of MPXV
Next Article in Special Issue
Epigenetic Regulation of β-Globin Genes and the Potential to Treat Hemoglobinopathies through Epigenome Editing
Previous Article in Journal
Salmonidae Genome: Features, Evolutionary and Phylogenetic Characteristics
Previous Article in Special Issue
Allele-Specific Disruption of a Common STAT3 Autosomal Dominant Allele Is Not Sufficient to Restore Downstream Signaling in Patient-Derived T Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

In Vivo Hematopoietic Stem Cell Genome Editing: Perspectives and Limitations

by
Nikoletta Psatha
1,
Kiriaki Paschoudi
1,2,
Anastasia Papadopoulou
2 and
Evangelia Yannaki
2,3,*
1
Department of Genetics, Development and Molecular Biology, School of Biology, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
2
Gene and Cell Therapy Center, Hematology Clinic, George Papanikolaou Hospital, Exokhi, 57010 Thessaloniki, Greece
3
Department of Hematology, School of Medicine, University of Washington, Seattle, WA 98195, USA
*
Author to whom correspondence should be addressed.
Genes 2022, 13(12), 2222; https://doi.org/10.3390/genes13122222
Submission received: 16 September 2022 / Revised: 11 November 2022 / Accepted: 22 November 2022 / Published: 27 November 2022
(This article belongs to the Special Issue Gene Editing for Therapy and Reverse Genetics of Blood Diseases)

Abstract

:
The tremendous evolution of genome-editing tools in the last two decades has provided innovative and effective approaches for gene therapy of congenital and acquired diseases. Zinc-finger nucleases (ZFNs), transcription activator- like effector nucleases (TALENs) and CRISPR-Cas9 have been already applied by ex vivo hematopoietic stem cell (HSC) gene therapy in genetic diseases (i.e., Hemoglobinopathies, Fanconi anemia and hereditary Immunodeficiencies) as well as infectious diseases (i.e., HIV), and the recent development of CRISPR-Cas9-based systems using base and prime editors as well as epigenome editors has provided safer tools for gene therapy. The ex vivo approach for gene addition or editing of HSCs, however, is complex, invasive, technically challenging, costly and not free of toxicity. In vivo gene addition or editing promise to transform gene therapy from a highly sophisticated strategy to a “user-friendly’ approach to eventually become a broadly available, highly accessible and potentially affordable treatment modality. In the present review article, based on the lessons gained by more than 3 decades of ex vivo HSC gene therapy, we discuss the concept, the tools, the progress made and the challenges to clinical translation of in vivo HSC gene editing.

Graphical Abstract

1. Introduction

For the last two decades, the development of an expanding set of genome editing tools is creating novel prospects in the field of gene therapy. The most extensively studied genome editing technologies include the Zinc-Finger Nucleases (ZFNs), the transcription activator-like effector nucleases (TALENs) and the Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR)-Cas9 nuclease (CRISPR-Cas9). A common feature of all these genome editing approaches is the precise modification of a specific DNA locus via artificial, programmable nucleases which produce double-strand breaks (DSBs) to a predetermined target genome site. The introduction of DSBs leads to the activation of endogenous DNA repair mechanisms: non-homologous end joining (NHEJ), homologous directed repair (HDR) or microhomology-mediated end joining (MMEJ). The most usually activated repair mechanism in human cells is NHEJ [1], which acts in an error-prone manner introducing small deletions and/or insertions (Indels) to the targeted genomic region, eventually disrupting the open reading frame. The HDR mechanism is more precise and requires cell cycling and a DNA donor-sequence as the repair template [2]. These site-specific modifications could lead to permanent genetic changes including the repair of a non-functional gene, the replacement of a missing/dysfunctional gene or an interference with gene expression.
Gene therapy and genome editing of hematopoietic cells have been explored mostly as an ex vivo approach. Ex vivo HSC genome editing is a personalized therapy requiring the production of a unique therapeutic product for each patient, generated from the patient’s own HSCs. The process involves patient mobilization and HSC collection, cell manipulation through electroporation or viral transduction, patient myeloablation and transplantation of the edited autologous graft. The multiple steps of this personalized therapeutic approach requiring specific infrastructure and transplantation expertise add to the complexity of the approach, thus increasing the possibility of an introduction of human error and limiting commercial access to the product due to exorbitant overall costs. Notwithstanding ex vivo gene therapy via gene addition or genome editing has, in recent years, led to unprecedented successes, it remains not broadly available and accessible to patients in need.
Towards overcoming current hurdles of ex vivo gene therapy/editing, in vivo approaches for HSC gene therapy have been proposed that would significantly simplify and expedite the delivery process, reduce cost and offer universal accessibility to patients. In vivo genome editing in non-hematopoietic tissues has been studied and well-argued in preclinical and clinical studies [3,4] whereas in vivo HSC genome editing is less well studied.
In this article, we briefly review the successes of ex vivo HSC genome editing and discuss the current status and the prospect of in vivo HSC gene editing as a therapeutic platform.

2. Genome Editing Modules and Ex Vivo HSC Gene Therapy

The development of genome editing techniques created new possibilities for gene correction and genomic modification of hematopoietic stem cells for several genetic and non-genetic disorders including β-hemoglobinopathies, primary immunodeficiencies (PIDs), congenital cytopenias and HIV.

3. ZFNs

Zinc-fingers nucleases (ZFNs) were amongst the first genome editing tools applied in HSCs. ZFNs are artificial recombinant nucleases composed of a Zn finger DNA-binding protein domain, fused with the endonuclease domain of the Fok1 restriction enzyme. The specificity of DNA targeting relies on a complex nucleotide-amino acid interaction code. The protein domain consists of three or more tandem zinc-finger individuals, each of which recognize three base pairs. Given that Fok1 dimerizes for activation, the binding of ZFNs pair to a specific DNA locus, resulting in the generation of a double-strand break (DSB) and the subsequent activation of endogenous DNA repair mechanisms [5,6].
Zinc-finger nucleases have been employed to target a variety of genomic loci as a curative means for both β-thalassemia and sickle cell disease. In combination with a single-stranded oligodeoxynucleotide (ssODN) as a donor template, ZFNs have been applied for an HDR-mediated correction of the SCD mutation in HSCs derived from SCD patients [7]. As co-inheritance of hereditary persistence of fetal hemoglobin (HPFH) with both β thalassemia and sickle cell disease (SCD) has been shown to ameliorate the severity of symptoms, ZFNs have been employed to reactivate the expression of the developmentally silenced fetal hemoglobin (HbF). Specifically, a ZFN-mediated inactivation of BCL11A, a major “silencer” of HbF expression, efficiently induced γ-globin expression in thalassemic HSCs [8,9,10]. The important identification of the erythroid enhancer of BCL11A, composed of three functional elements in +55 kb, +58 kb and +62 kb, made possible the functional disruption of BCL11A expression in the erythroid lineage [11]. In the first clinical trial [NCT03432364; ST-400-01] by Sangamo Therapeutics and Sanofi, rapid hematologic reconstitution, HbF elevation and a persistence of editing rate was observed six months after transplantation in thalassemic patients, while SCD patients remained symptom-free for up to 52 weeks post CD34+-edited cell transplantation [12].
In immunodeficiencies, both inherited and acquired, ZFNs have been applied in several gene editing approaches. X-linked Severe Combined Immunodeficiency (X-SCID), characterized by mutations in the IL2RG gene and resulting in impaired humoral and cell immunity [13], was effectively corrected by the HDR-mediated editing of IL2RG via ZFNs, for both in patient-derived HSCs and xenotransplantation mouse models [14,15,16]. Wiskott-Aldrich syndrome (WAS), another rare X-linked congenital syndrome caused by mutations in the WAS gene and playing an important role in actin cytoskeleton remodeling specifically in HSCs [17], was also targeted by ZFN gene editing. ZFN-mediated introduction of the wild type WAS exons 2-12 in the first intron of the WAS gene resulted in a restoration of the protein expression (WASp) in patient-derived iPSCs, leading to an efficient differentiation of edited iPSCs into functional NK and T-cells [18]. In acquired immunodeficiencies, such as Human Immunodeficiency Virus (HIV) infection, the ZFN-mediated disruption of the CCR5 locus—a principal receptor for HIV entrance into host cells—in CD4+ T-cells or in CD34+ cells significantly reduced the viral replication in non-human primates [19].

4. TALENS

Transcription activator-like effector (TALE) proteins are derived from the phytopathogenic bacterial genus Xanthomonas and are characterized by their DNA-binding ability. A wide range of effective domains such as nucleases, transcription activators or suppressors and site-specific recombinases can be fused with TALE proteins, supporting different genomic manipulations. Of these, the most common combination consists of the fusion of TALE-binding proteins with Fok1 nuclease to generate TALE nucleases (TALENs). Thus, TALENs comprise four functional domains: a nuclear localization signal (NLS), an acidic domain for target gene transcription activation, a central DNA-binding domain of 12-28 amino acid tandem repeats and a Fok1 nuclease [20]. Every single repeat within the DNA-binding domain consists of 30-35 amino acids and recognizes a single nucleotide respectively [21,22]. Given the fact that Fok1 nuclease has to be dimerized in order to generate a cleavage on both of the DNA strands, TALEN modules must be designed in pairs to bind opposite DNA target sequences [23].
TALENs have been used in β-hemoglobinopathies for the reactivation of γ-globin genes in a similar fashion to ZFNs. Specifically, 13nt deletion in HBG1 and HBG2 promoters via TALENs mimics a well-known, naturally occurring HPFH mutation resulting in HbF reactivation [24]. In addition, TALENs have been shown to inactivate the expression of BCL11A in CD34+ cells of non-human primates [25] or to correct the HBB IVS-110(G>A) mutation in human erythroblasts derived from IVS-110(G>A)-homozygous patients [26]. A targeted correction of the IVS2-654 C > T HBB gene mutation via homologous recombination using TALENs was demonstrated in β-thalassemia-derived iPSCs [27] and in a β-thalassemia mouse model [28]. Immunodeficiencies have also been targeted with the TALEN platform. Specifically, Menon et al. efficiently utilized TALENs in combination with a donor sequence to correct the c.468+3A>C mutation of the IL2-Rγ gene in patient-derived iPSCs [29], while Cellectis applied the TALEN® technology to precisely correct RAG-1 mutations in preclinical studies [30]. In HIV, TALEN-mediated disruption of CCR5 in human CD4+ T-cells presented high specificity and efficacy against the infection [31,32].

5. CRISPR-Cas9

The discovery of the Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR)-associated nuclease Cas9 (CRISPR-Cas9) system as a feature of adaptive immunity in bacteria and archeas led to a tremendous progress in genome editing. Τhe CRISPR-Cas9 system consists of a guide RNA (gRNA) complementary to a desired DNA sequence and a Cas9 nuclease [33,34]. After the recognition of the protospacer adjacent motif (PAM) at the 3′ end of the targeted DNA locus and the hybridization of gRNA with DNA sequence, the Cas9 nuclease is tied to the DNA-gRNA complex and generates a DSB on the specific locus which activates the endogenous DNA repair pathways [33,35].
The development of the CRISPR-Cas9 system has played a crucial role in the improvement of genome editing procedures for HSC gene therapy and the expansion of their potential application. In β-hemoglobinopathies, a generation of HPFH-associated mutations in the HBG promoters via a CRISPR-Cas9-mediated disruption of HbF suppressors-binding domains effectively increased HbF levels in both SCD and thalassemic patient-derived HPSCs, while it ameliorated the disease phenotype in xenografts [36,37,38]. Alternatively, the CRISPR-Cas9-induced disruption of HbF key regulators and their regulatory sequences (i.e., BCL11A, KLF1 and ZBTB7A), an approach already successful with ZFNs, was thoroughly investigated [39,40,41,42]. Several groups worked on the erythroid specific BCL11A disruption by targeting the aforementioned erythroid enhancer in the BCL11A locus, either individually or in combination with BCL11A-binding sites within the HBG1/HBG2 promoters [36,43,44,45]. In 2018, VERTEX Pharmaceuticals, in collaboration with CRISPR-Therapeutics, initiated two clinical trials for Transfusion-Dependent β-Thalassemia (TDT) [NCT03655678; CTX001-111] and Sickle-cell disease [NCT03745287; CTX001-121] based on ex vivo genome editing of BCL11a erythroid enhancer via CRISPR-Cas9 in CD34+ HSCs. Recently, disclosed results from the two studies at the EHA 2022 meeting demonstrated that 42 of 44 TDT-patients stopped RBC transfusions for up to 36 months, maintaining after month 3 mean total Hb levels >11 g/dL, of which more than 90% was HbF. SCD patients (n = 31) remained severe vaso-occlusive crises (VOCs)-free for a maximum duration of 32 months, having stable levels of approximately 40% HbF after month 4 [46]. Furthermore, Novartis has generated two gene editing products, OTQ923 and HIX763 [NCT04443907] for SCD, that reduce the activity of BCL11A thus resulting in HbF reactivation. Finally, two additional clinical trials, supported by Graphite Bio, Inc [NCT04819841; GPH101-001; CEDAR] and Editas Medicine, [Inc NCT04853576; EM-SCD-301-001] explored the CRISPR-Cas9 system for ex vivo editing of CD34+cells from SCD patients [47].
Fanconi Anemia (FA) has been a continuing challenge as a target disease for gene therapy over the years. A promising approach for gene therapy of Fanconi Anemia (FA) is the generation of compensatory mutations in the mutated FANCA gene with the CRISPR-Cas9 technology in order to restore its functions. Indeed, editing of the FANCA locus in HSPCs derived from FA patients through the introduction of compensatory mutations, led to the correction of cell phenotypes without affecting their differentiation and self-renewal capacity, as their transplantation into NSG recipients resulted in increased engraftment levels and a demonstrated in vivo proliferative advantage of edited cells [48].
The Wiskott-Aldrich immunodeficiency is caused by >300 different mutations of the WAS gene. Correction of each of these mutations with gene editing, although theoretically feasible, would have been impractical and expensive. Using the gene editing platform to knock-in a therapeutic WAS cDNA in frame with its endogenous translation starting codons in patient-derived HSPCs may ensure the correction of all known disease-causing mutations. Indeed, such HDR-mediated replacement of a mutated gene by a wild-type WAS cDNA via CRISPR-Cas9 led to the restoration of WAS protein levels and the improvement of T-cell functions [49].
As with the other gene editing platforms, the CRISPR-Cas9 system has also been extensively applied in HIV gene therapy. CXCR4, as well as CCR5, were efficiently disrupted in primary CD4+T-cells via CRISPR-Cas9 thus conferring resistance to HIV tropism [50,51]. Remarkably, the transplantation of allogeneic, CRISPR-edited, CCR5-ablated HPSCs into a patient with HIV-1 infection and acute lymphoblastic leukemia resulted in long-term engraftment of genetically modified HSPCs, yet with only 5% CCR5 disruption rates in lymphocytes [52].

6. Double-Strand Break-Free Gene Editing

The evolution of CRISPR-Cas9 technology resulted in a tremendous optimization of genome editing approaches. However, concerns and limitations still exist; much attention has been paid to the unintended, “off-target” genotoxicity, which has now been well described while much less is known about “on-target” detrimental consequences arising from DSBs.
Off-targeting occurs due to nonspecific CRISPR/Cas-induced DNA cleavage at sites other than the actual target but with substantial sequence similarity to the intended target, thus potentially providing deleterious effects. In addition, DSB-induced activation of p53 with subsequent DNA damage response (DDR) and cell-cycle arrest has been described to decrease the HSPCs function and impair HDR, even though the delivery of Cas9 ribonucleoproteins has a limited lifetime within cells [53]. Transient p53 inhibition has been shown to enhance HDR efficiency, resulting in polyclonal reconstitution of the engrafted HSPCs [54,55].
In addition to the off-target effects, “on target” consequences include rearrangements and several kilo- to mega-bases telomeric deletions, ablation of entire chromosomes or the recently described chromothripsis, whereby massive genomic rearrangements occur in a one-off cellular crisis and can potentially generate a malignant outcome [56,57,58,59].
Collateral damage by gene editing may be permanent and irreversible. The accumulating evidence on the undesired either off-target or on-target consequences in clinically relevant primary cells warrants an optimization of the current editing systems and methods for genome-wide profiling of off-target effects and for decreasing or avoiding CRISPR off-target activity [60].

6.1. Base Editors

Gene editing strategies that do not generate DSBs, including base or prime editors (BE, PE), are considered substantially safer over traditional gene editing approaches as they overcome the DSB-associated deleterious effects in genome integrity [61]. Base editing precisely generates targeted point mutations without generating DSBs or requiring donor DNA templates and activation of the endogenous HDR mechanism [62].
Cytosine base editors (CBEs) consist of a cytidine deaminase fused to a mutated form of Cas9 from Streptococcus pyrogenes which is unable to generate DSBs (dead Cas9; dCas9). After binding to its target site, dCas9 performs local denaturation of the DNA complex to generate single strand DNA chains. Cytidine deaminase converts the desired cytosine to uracil in the strand which is not paired with guide RNA. Cell replication machinery recognizes uracil as a thymidine resulting in a C-G to A-T transition [63,64,65,66].
Adenosine base editors (ABEs). In contrast to DNA cytidine deaminases, there is not an enzyme catalyzing the deamination of adenine to inosine in nature. Up to date, all reported enzymatic inversions of adenine to inosine occur on free adenine, free adenosine, adenosine in RNA or adenosine in mis-paired RNA-DNA heteroduplexes [64,67]. To overcome this obstacle, Gaudelli et al. created an engineered enzyme based on a tRNA adenosine deaminase enzyme, TadA, originating from Escherichia coli. The recognition of adenosine by TadA results in hydrolytic deamination of adenosine. The remaining inosine is recognized by a cellular repair mechanism as a cytosine, and the intermediate T-I base pair is replaced by a G-C base pair [68].
Efforts to treat β-hemoglobinopathies with base editing have shown that targeting the +58 BCL11A erythroid enhancer with an A3A(N57Q)-BE3 efficiently converted the desired G-C to an A-T which led to the suppression of BCL11A expression and increased the levels of HbF [69]. Wang et al. used a hAPOBEC3A-Cas9n (hA3A-BE3) to introduce single nucleotide substitution at -115C and -114C in HBG promoter, mimicking HPFH mutations and increasing γ-globin expression from ~6.8% to ~44.2% [70]. Recently, an ABE- or CBE-mediated introduction of clustering mutations ~200 bp upstream of HBG1/2 genes effectively reactivated γ-globin expression in SCD HSPCs via the disruption of LRF binding or the induction of KLF1 recruitment [71].
SCD is caused by a single base-pair point mutation (GAG to GTG) due to the replacement of glutamine acid (GAG) by valine (GTG), which could be an ideal base editor target. Nevertheless there is currently no base editor able to convert T to A. However, Newby et al. were able to generate an adenine base editor converting the GTG codon to GCG, leading to the expression of a non-pathogenic variant known as Hb-Makassar (HBBG) [72]. C. Li. et al. created an ABE in order to generate -113A>G mutation in the HBG promoter mimicking an HPFH mutation. In vivo editing of the promoter in β-YAC/CD46tg mice resulted in a 20% conversion rate in HSPCs and >40% γ-globin expression in peripheral RBCs [73]. The HBB -28A>G mutation is one of the most frequently detected mutations in β-thalassemia patients in China and East Asia preventing the transcription of the HBB gene [74]. Two base editor variants, the eA3A-BE and eA3A(N57Q)-BE3, were employed in erythroid precursor cells derived from a compound heterozygous thalassemia patient (4bp-deletion in exon 1 of one HBB allele and HBB -28 A>G mutation in the other), and they effectively generated the C>G substitution in HBB-28. Importantly, eA3A-BE or eA3A(N57Q)-BE3 editing increased HBB expression by 2.6 and 4.0-fold, respectively, compared to control samples [75].
The base editing technology has been applied in other disorders in addition to hemoglobinopathies. In particular, simultaneous disruption of the HIV receptors CXCR4 and CCR5, either in primary human T-cells via cytosine base editors or in primary T-cells, and CD34+ HSPCs via adenine editors efficiently disrupted CXCR4 and CCR5 expression thus protecting from CXCR4- and CCR5-tropic viral infections [76].
Adenine base editors have also been applied for the targeting of the FA-55 and FA-75 mutations in the FANCA gene resulting in the restoration of expression and phenotypic correction in HSPCs derived from a Fanconi anemia patient [77]. Finally, ABEs have been recently used for the targeted correction of CD3D C202T, a mutation causing CD3δ severe combined immunodeficiency (SCID) in Jurkat T-cells and in CD34+ HSPCs leading to a more efficient CD3 repair compared to a CRISPR-Cas9 correction via homologous recombination [78].

6.2. Prime Editing

Despite the enormous progress in precise genome editing with the discovery of base editors, the inability to install all possible combinations of bases substitutions represents a challenge that needs to be addressed. Recently, Prime Editors (PE), a versatile genome editing platform, has been generated [62]. The major advantage of prime editors is their ability to accomplish all twelve types of base pair conversions alone or in combination with the installation of small deletions/insertions in DNA, without the generation of DSBs. The prime editors (PEs) consist of a Cas9 nickase fused to an engineered reverse transcriptase (RT) [79,80]. The PEs use a prime editing guide RNA (pegRNA) which contains a sequence complementary to the desired DNA sequence, a prime editor binding site and the sequence that will be introduced to the genome after RT activation. Because PEs incorporate the edit in one of the two DNA strands, in order to manipulate the DNA repair mechanism to use the edited strand as a template for repairing the non-edited strand, an additional gRNA is used for creating a nick in the non-edited strand away from the initial nick [80,81].

6.3. Epigenome Editing

The epigenome is shaped by chemical compounds that modify or mark the genome being inheritable during cell division, albeit not part of the DNA itself. Thus, epigenome editing refers to the modification of the epigenome using engineered tools aiming to modulate the chemical state of DNA structure and function, representing an alternative way through which a cell’s phenotype and/or function can be altered without modifying the underlying DNA sequence [82].
In the last years, several studies have revealed the impact of epigenome alterations on gene expression and the development of genetic disorders and cancer as a result of the above. There are various epigenomic modifications that tightly control cellular processes [83]. DNA methylation is a major epigenetic modification known to cause a silencing of gene expression [84]. Another regulatory process is histone modification, which is induced by methylation and acetylation on histones located in the vicinity of promoters or enhancers. Histone methylation/demethylation are orchestrated by histone methyltransferases/demethylases, respectively, and histone acetylation/deacetylation are catalyzed by histone acetyltransferases (HAT)/deacetylases (HDAC), respectively [85,86].
The combination of the hitherto acquired knowledge in the field of epigenomic modifications with the innovative genome editing approaches has led to the generation of epigenome editors (epi-editors) [87]. The first epigenome editors were generated after fusion of the catalytic domains of enzymes such as DNMT3 to the catalytic residues of programmable DNA binding molecules (TALE, ZFP) [88,89]. Later, CRISPR epigenome editors were designed, consisting of a catalytically inactivated “dead” Cas9 (dCas9) fused or non-covalently bound to the catalytic domain of epigenetic effectors such as DNMT or TET enzymes, HATs or HDACs to activate or repress gene expression. A gRNA complementary to the target DNA sequence navigates the CRISPR-Cas9 epigenome editor to the target site, usually a promoter or distal cis-regulatory sequence [90,91]. Other epigenome editing tools transiently expressing transcriptional repressors, such as DNMT3A or a combination of the DNMT3a and KRAB domains to target the regulatory sequences of a gene of interest have been efficiently established. Specifically, inheritable targeted epigenome silencing was achieved in normal T-lymphocytes by inducing repressive histone marks and de novo DNA methylation [92]. Another system relying on the recruitment of Cas9 and transcriptional activation complexes to target loci by modified single guide RNAs has been developed to activate silenced endogenous target genes through trans-epigenetic editing. Proof-of-concept preclinical studies have shown that in vivo CRISPR/Cas9-mediated target gene activation (CRISPRa) improved phenotypes in relevant mouse disease models [93]. Moreno et al. recently coupled dCas9 with several transcriptional regulators, achieving multiplex targeting via single or dual-gRNA delivery that resulted in a high level of in vivo transcriptional repression (up to 80%) and transcriptional activation (up to 6-fold increase). This multiplex gene activation and/or repression approach could be beneficial for complex diseases that have multiple genomic loci involved [94]. CRISPRoff and CRISPRon are two technologies developed for programmable writing and erasing epigenetic memories; transient expression of CRISPRoff writes a robust, specific, and multiplexable gene-silencing program that is memorized by human stem cells through cell division and differentiation into neurons and can be rapidly reversed by CRISPRon [95].
Undoubtedly, epigenome editing-mediated transcriptional control may provide a platform for powerful and highly personalized therapeutics. Since this system does not rely on DNA breaks and genomic sequence modification, it is perceived to be reversible and less permanent, thus intrinsically safer. Nevertheless, the knowledge regarding the epigenetic editing effect in primary cells is limited, and basic biological questions still need to be addressed towards a safe translation to the clinic.

6.4. RNA Editing

The discovery of the Cas13 nucleases family, comprising 6 subtypes (a, b, c, d, X, Y) and having exclusively single-stranded RNA-targeting properties, further expands the range of genome editing approaches [96,97]. The unique feature of Cas13 nuclease is the RNase activity after an RNA-RNA recognition. The CRISPR-Cas13 system includes a CRISPR- RNA (crRNA) identifying a specific sequence on the target RNA. After the hybridization of crRNA with the targeted RNA sequence, Cas13 binds to the complex and cleaves in a specific position [98,99]. This approach leads to knock down of a specific gene without the interruption of a DNA sequence. Moreover, the substitution of Cas13 by catalytically inactive Cas13 (dCas13) can transform the CRIPR-Cas13 system into a programmable RNA binding tool.
The fusion of dCas13b with catalytic domains of RNA-modifying enzymes, such as the adenosine deaminase domain of adenosine deaminase acting on the RNA-2 (ADAR2) proteins, led to the conversion of adenosine to inosine, yet with a substantial number of off-target RNA-editing effects. The latter was addressed by further engineering of the system to create an ADAR2 variant capable of precise, efficient and highly-specific editing when fused to dCas13b [100,101]. Recently, Cas13 was harnessed to target known RNA viruses including lymphocytic choriomeningitis virus (LCMV), influenza A virus (IAV), vesicular stomatitis virus (VSV) and severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) [102,103]. Yin et al. showed that expression of CRISPR-Cas13a in HIV-effected cells successfully destroys viral RNA and prevents HIV reinfection in HEK293T human cell line, presenting an alternative potent approach for RNA editing in HIV infected T-cells [104].

7. Delivery Tools for In Vivo Gene Therapy

Most current HSC gene-addition and -editing clinical trials involve ex vivo manufacturing in which cells are harvested from a patient’s body by leukapheresis, engineered outside the body and then reintroduced into the patient, usually after a myeloablative conditioning. The ex vivo approach is highly sophisticated and cost-intensive while it requires transplantation expertise and infrastructure. In vivo gene therapy, namely the direct injection of the therapeutic vector into the patient in the absence of conditioning, has been extensively used for gene therapy of target tissues not easily accessible, such as liver, muscle, lung epithelium and neurons.
The bone marrow stroma generates a physical barrier to the transduction of HSCs with IV-injected gene transfer vectors, and this has been a major challenge for in vivo HSC gene therapy. Intraosseous (IO) delivery of hematopoietic cells has been explored as a means to mitigate the loss of cells given IV to high blood volume organs, such as the liver and the lungs. Although sustained transduction of hematopoietic stem/progenitor cells has been shown after lentiviral vector IO delivery post reduced intensity conditioning, the approach is technically challenging and invasive whereas it requires further optimization [105].
Based on the lessons gained during the evolution of gene-addition therapy especially as regards the in vivo gene delivery to non-hematopoietic tissues, spanning more than 3 decades of research and clinical translation, we will discuss the recent developments and challenges of in vivo gene therapy of HSCs.

7.1. Lentiviral Vectors

Lentiviruses, such as the human immunodeficiency virus (HIV), are RNA viruses of the Retroviridae family. HIV-derived lentiviral vectors (LVs) have been the gold standard for transgene delivery into HSCs due to their ability to transduce non-dividing cells and their safer genome integration pattern over γ-retroviral vectors [106,107] while they represent, up to now, the preferred tools for ex vivo gene correction. Self-inactivating lentiviral vectors with an enhanced safety profile have been used in numerous clinical trials for ex vivo HSC gene addition/editing, providing remarkable therapeutic benefits in patients with severe inherited blood disorders such as the Wiskott–Aldrich syndrome (WAS) [108,109], X-linked and ADA severe combined immunodeficiency (SCID-X1, ADA-SCID) [110,111], β-hemoglobinopathies [112] and neurodegenerative storage diseases (adrenoleukodystrophy and metachromatic leukodystrophy) [113,114].
For tissues with limited accessibility, such as the liver, in vivo lentivirus-mediated gene therapy specifically targeting the liver had been suggested from Luigi Naldini and Didier Trono early on, in 1996 [115]. However, LV-induced immune responses, by the in vivo administration, consisted of a major limitation leading to the reduction of transduction efficiency, the rejection of transduced cells and an overall inhibition of the therapeutic effect [116]. This effect was caused by the incorporation of the packaging-cell derived polymorphin class-I major histocompatibility complexes (MHC-I) in the virus’ surface. Packaging lentivirus in B2M-deficient cells can significantly reduce immune responses to the virus while not affecting viral titers and virus incorporation into the target cells [117].
LVs are modified to bear different viral envelopes (pseudotyping), most commonly the vesicular stomatitis virus glycoprotein (VSV-G), to alter and improve their tropism for different cell types. However, the inactivation of VSV-G pseudotyped particles by the human serum [118], as well as the lack of the LDL receptor in quiescent HSCs [119], represent obstacles for in vivo delivery of LVs to HSCs. Engineering the VSV-G glucoprotein to generate human serum-resistant and thermostable VSV-G variants [120], employing different envelops to pseudotype LVs (e.g., H/F-LVs, cocal LVs, measles and BaEV-LVs) [121,122,123] or using different retroviral vectors such as foamy viruses [124] represent approaches shown to achieve high-level transduction of unstimulated HSCs that was maintained in all hematopoietic lineages and in secondary-recipient NSG mice [125] or to successfully correct disease phenotypes by in vivo gene delivery (X-SCID) [121,122].
From a different perspective, the ability of LVs to accommodate relatively large transgenes, with a packaging size of about 10 kb, renders them ideal for the delivery of several genome editing tools in vitro or ex vivo. All the previously described editing systems can be comfortably fitted within the LVs with ample room for selection cassettes, donor templates or suicide genes. However, the stable, life-long expression of a DNA-editing protein delivered by the integrating to the genome LVs would pose a genotoxicity threat to the patients and possibly lead to intolerable toxicity. To overcome this issue, LVs depleted of their ability to integrate into the host genome (Integration deficient lentiviral vectors-IDLVs) or to reverse transcribe their transgene (non-reverse transcribable lentiviral vectors-NRTLVs) have emerged as a promising possibility [126,127,128].

7.2. Adeno-Associated Viral Vectors

Adeno-associated viruses (AAVs) are single-stranded, replication-deficient DNA viruses with a compact 4.7-kb genome flanked by palindromic inverted terminal repeats (ITRs). The AAVs represent a prominent tool for in vivo gene therapy for many diseases with >280 registered clinical trials world-wide [47] (ClinicalTrials.gov (accessed on 12 September 2022)), and several AAV-based therapeutic products (i.e., Glybera, Luxturna and Zolgensma) [129,130,131] have been approved by regulatory authorities. AAVs, like LVs, are capable of transducing both dividing and non-dividing cells [132], a requirement for in vivo HSC transduction. Their main advantage is that they are derived from viruses that have been evolutionary selected for transducing human cells, at the expense, however, of a human immune system response against these viruses and the vectors derived from these viruses, resulting in immune-mediated rejection after in vivo delivery [133].
Out of the wide AAV-serotype ranges available, AAV2 has been the first to be used in HSCs with positive outcomes [134,135,136], even though variable early results are attributed to differences in viral titers and donor variation [137,138,139]. In 2013, AAV6 was established as the optimal serotype to be used in human HSCs both in vitro and in vivo [140]. Subsequently, several groups have attempted to use AAV6 for the ex vivo delivery of a donor template for HDR-mediated targeting integration in HSCs [36,49,141,142,143,144], reaching an editing efficiency of 90% in vitro [145]. Even though delivery of a donor template through AAV6 has been largely effective, the use of AAVs as multiplex vehicles to deliver more complex and larger cassettes such as genome editing modules is challenging, mostly due to their low packaging ability.
ZFNs, with their monomer cDNA size approximately at 1kb, have been easily incorporated in AAVs since the early 2010s, either alone or in combination with a donor template for targeted integration [146,147]. However, encompassing larger sequences such as TALENs (3 kb/monomer), SpCas9 (4.2 kb), base (>5 kb) and prime (>6 kb) editors might prove challenging and require splitting the transgene cassettes in two separate vectors, a rather undesirable approach in an in vivo editing setting. As an alternative to AAV-spCas9 delivery, smaller Cas9 or Cas13 orthologs, such as sauCas9 (3.2 kb) or Cas13d [148,149] and CasRx (~3 and <4.3 kb, respectively), have been effectively incorporated in AAVs for DNA, RNA or epigenome editing [99,150,151,152].
Another major obstacle for the use of AAVs in vivo is the expected immune response and presence of neutralizing antibodies (NAb), as the wild-type AAVs infect humans at a young age without, however, developing any known pathology [153]. Studies in both small and large animal models described a prevalent serum neutralization of AAV6 transduction [154] while similar findings in humans for almost all serotypes have prompted the establishment of an NAb cut-off exclusion criterion for clinical trial enrollment [155,156,157,158]. Transduction inhibition due to neutralization, among other factors, created the need for higher viral doses which, in turn, were associated with acute liver toxicity. The latter, unfortunately, has led to the untimely death of four children and one young man in clinical trials of AAV8- and AAV-9-mediated correction of X-linked myotubular myopathy and Duchenne’s muscular dystrophy (XLMTM), respectively [159,160]. Acute liver failure was also the cause of death in two kids with spinal muscular dystrophy, 5-6 weeks after receiving the commercially approved AAV-9-based gene therapy product, Zolgensma [161].
Additional hurdles for AAV-mediated in vivo HSC gene therapy include the limited knowledge of AAV tropism to HSCs and their mainly episomatic cell localization, which, due to their unsuitability for multiplex manufacturing, rather precludes their use for targeting HSCs.

7.3. Adenoviral Vectors (Ads)

Adenovirus is a ~100 nm, non-enveloped, double-stranded DNA virus first isolated in 1953 [162]. Their ability to efficiently transduce a wide range of dividing and non-dividing cells, along with their well-known genome sequence, at early times led to the first in vivo gene delivery approach in an animal model [163]. The first human studies, however, demonstrated strong innate, humoral and cellular immune responses elicited by the Adenoviral vectors [164,165] which, in some cases, led to serious adverse events such as cytokine storm resulting in a patient’s death in 1999 [166]. Several strategies to circumvent these acute immune responses have been explored including the administration of immunosuppressive agents or monoclonal antibodies blocking cytotoxic T lymphocyte (CTL) immune responses [167,168,169,170,171,172].
To date, over 57 serotypes (Ad1-57) have been identified and divided in groups A-G. Of them, Group C Ad5 is the most well characterized and broadly tropic, albeit not towards HSCs. In contrast, Group B Adenoviruses, including Ad3, Ad11, Ad35 and Ad50 can transduce human HSCs via either desmoglein 2 (DSG2) [173] or the surface CD46 protein [174]. To harness the properties of Ad5 and the enhanced HSC transduction capabilities of group-B Ads, fiber-chimeric vectors containing B fibers on an Ad5 capsid were developed, and vectors incorporating the fiber of the CD46-tropic Ad35 (Ad5/35) were shown to efficiently transduce human HSCs [175,176] without causing liver toxicity after IV injection, as opposed to Ad5 vectors [177].
In addition to immunomodulating strategies to minimize the AdV vector-elicited immune responses, the development of 3rd generation “gutless” or Helper-Dependent (HDAd) vectors, deprived of all viral sequences, increased both the safety profile of the vectors and the efficiency and duration of the therapeutic effect in vivo [178,179]. In contrast to AAV, the large transgene capacity (>30 kb) of HDAds and their low manufacturing cost makes them ideal for the accommodation of any genome-editing modality or possible combinations.
Ads do not integrate their genetic material into the host cells’ genome. This feature would have precluded these vectors for HSC gene therapy which requires lifelong gene correction; however, recently, hybrid Hd5/35 vectors exploiting transposon-based transgene integration systems (Sleeping Beauty, piggyBac, and Tol2) [180] have successfully integrated transgenes into the genome [181,182].
A novel, simplified and minimally invasive platform for in vivo HSPC gene therapy using a hybrid vector system comprising a HDAd5/35++ vector with increased CD46 affinity for transgene delivery to primitive HSCs and a hyperactive Sleeping Beauty transposase (SB100X) for transgene integration has been developed by the lab of A. Lieber [183]. This platform consists of a mobilization round with G-CSF + Plerixafor followed by injection of the vector when the circulating HSPCs are in their highest concentration. Transduced cells home back to the bone marrow where they persist and stably express the transgene [182]. With this approach, which could be also applicable using other delivery methods, choosing the optimal mobilization scheme relative to the patient’s disease background comprises an important factor. For example, the G-CSF + Plerixafor combination is now considered to be superior to the standard G-CSF mobilization β-in thalassemia [184,185,186,187] and Fanconi Anemia [188], while G-CSF is contraindicated in patients with sickle cell disease and for SCD gene therapy, Plerixafor-mobilized HSCs are harvested [189,190].
The HDAd5/35++ in vivo HSC targeting platform has broad implications for gene addition and gene-editing therapy of inherited or acquired diseases that require high levels of therapeutic proteins in the blood circulation. Indeed, up to date, this approach has been shown to safely and efficiently transduce primitive HSCs or/and ameliorate or correct disease phenotypes either using stably expressing gene-addition systems for β-thalassemia, SCD, hemophilia [191], X-SCID [192] and SARS-Cov2 [182,191,193,194,195,196,197] or precision editing mainly for β-hemoglobinopathies and HIV [44,73,193,198,199,200] in human HSCs, mouse models and non-human primates [201]. This extensive work demonstrated that HDAds can comfortably carry several genome-editing systems in HSCs, such as ZFN dimers [198], CRISPR/Cas9 with one or more gRNAs [44] and base editors [73].

7.4. Non-Viral Transfer

Non-viral transfer offers an alternative delivery tool to overcome limitations associated with the viral vectors including the reduced packaging capacity (AAV) [202], the pre-existing humoral and cell immunity against certain viral serotypes leading to virus neutralization and reduction of in vivo transduction efficiency [203] and the increased risk of off-target effects and insertional mutagenesis with the prolonged presence of genome editing tools into non-dividing cells [204,205].
The non-viral gene-delivery methods are divided into two different categories: the physical methods such as electroporation, microinjection, hydrodynamic delivery and the chemical methods [206]. Herein, we focus on chemical methods which can be translatable in vivo. This category includes organic nanoparticles composed of lipid or peptide-based materials and natural or synthetic polymers. In some cases, inorganic agents such as calcium phosphate or metals are used for nanoparticle formation [207,208,209,210]. Until now, non-viral chitosan, Poly-(lactic-co-glycolic) acid (PLGA) and cationic lipid nanoparticles have been successfully applied in experimental protocols [211] as well as in gene therapy trials for cystic fibrosis [212,213]. Furthermore, organic nanoparticles have been used for targeting liver cells [214,215]. In a clinical trial for Transthyretin amyloidosis gene therapy [NCT04601051], lipid nanoparticles have efficiently been used for in vivo delivery of a CRISPR-Cas9 system resulting in transthyretin (TTR) protein disruption [216].
Although the non-viral delivery methods in HSC gene therapy are not widely applied, the replacement of viral platforms with chemical tools could overcome their associated side effects and decrease the cost of the procedure. PLGA nanoparticles, loaded with triplex-forming peptide nucleic acids (PNAs) and single-stranded donor DNA molecules introducing site-specific repair and recombination, were used to specifically modify either the CCR5 gene or the β- or γ-globin gene in relevant mouse models in order to prevent HIV infection or correct the βIVS2-654 mutation, respectively [217,218]. Recently, Cruz et al. exploited PLGA-nanoparticles for CRISPR-Cas9 delivery into primary erythroblasts and human CD34+ HSCs to reactivate γ-globin expression [219]. The development of layer-by-layer (LbL) nanoparticles, which consist of a nucleic acid core, a negative charged layer and a non-degradable synthetic peptide, enhanced the delivery efficiency and made possible the targeted transport into HSCs. Specifically, LbL nanoparticles containing an outer layer of anti-CD117/c-kit antibodies attached to hyaluronic acid efficiently targeted in vivo mobilized HSCs in a mouse model [220].

8. Current Issues and Considerations for In Vivo Gene Therapy

The in vivo gene therapy has clearly several advantages as compared to the ex vivo approach. It is minimally invasive and simplified, thus abrogating the need for leukapheresis, myeloablative conditioning with chemotherapy and transplantation expertise as well as the barrier of limited patient accessibility to ex vivo gene therapy products due to their exorbitant costs. Moreover, by skipping the HSC ex vivo manipulation, the impaired homing/engraftment of transduced cells is avoided and all HSCs, including the “true” stem cells that could have been missed by isolating the HSCs on the basis of the “conventional” CD34+cell marker, can be targeted.
Undoubtedly, the future of in vivo HSC genome modification seems prosperous. However, there are still limitations that scientists should be able to overcome and explore before attempting this approach en masse. Vectors for in vivo HSC gene-addition therapy need to integrate into the genome and for in vivo gene editing to specifically and precisely bind to HSCs without off-target delivery. The primarily quiescent nature of HSCs makes in vivo gene targeting highly challenging, especially when nuclease-mediated HDR, restricted to the G2/S cell cycle phase, is considered and, in this context, NHEJ, base or prime editors should be more appropriate.
The optimal type of editor to be selected is crucial and reasonably high editing efficiency is a major prerequisite. However, efficiency must be coupled with retaining the modified cell’s fitness and function, minimal—if any—off-target effects and, of major importance, a low immunogenicity profile. In vivo gene therapy faces challenges from both the innate and adaptive immunity; pre-existing immunity and immunotoxicity represent a significant barrier for in vivo delivery, as several of the editing modules are of microbial nature and peptides of the programmable nucleases might be presented by Major Histocompatibility Complex (MHC) Class I molecules [221,222,223,224,225]. Pre-existing B- and T-cell responses against capsid proteins of Ad and AAV vectors can neutralize the vector before it transduces the HSCs and also prevent re-administration. Neutralizing antibodies against AAV1,2 and 6 and Ad5 [226] are prevalent in serum, whereas Ad35 is a rare human serotype and thereby Ad35 vectors evoke only mild host immune responses and contribute to prolonged gene expression [227]. Transient immunosuppression using corticosteroids with or without anti-IL-6R (tacrolimus) has been effective to confront humoral immunity in in vivo liver-targeted gene therapy for hemophilia B using AAV8 or AAV3 vectors [228,229]. Mobilization before IV vector administration may further increase cytokine release in response to an innate immune reaction. Nevertheless, pretreatment with dexamethazone, tacrolimus and anti-IL-1 to suppress the inflammatory cytokine storm allowed for safe in vivo gene editing in mobilized NHPs with the HdAd5/35 vector system [201].
Off-target effects might also differ significantly between nucleases and be also specific to the target locus. Detection of such off-target events is especially challenging with in vivo editing. It might be feasible though by an unbiased high-sensitivity deep whole genome sequence [230,231]. Recently, a dual step assay where the genomic loci harboring off-target events are identified in vitro and, subsequently, the same sequences are assayed in vivo has been described [232]. Additional methods for on- and off-target nuclease cleavage detection have been extensively reviewed by Andrew Atkins et al., 2021 [233]. In addition to the introduction of mutations in unpredicted or undesired genomic locations, double-strand brakes by programmable nucleases and especially CRISPR/Cas9 have been shown to cause large deletions and genomic rearrangements [234], chromothripsis [56] and aneuploidy [235].
Methods to mitigate current limitations are compulsory before broadly transferring these therapeutic approaches in the clinic. The continuously increasing developments in the field and the discovery of new, optimized and safer genome-editing modules and platforms may enable in vivo gene addition or editing and become clinically applicable, widespread accessible and an affordable treatment for all patients.

Author Contributions

Conceptualization, N.P. and E.Y.; writing—review and editing, N.P., K.P. and A.P.; supervision, E.Y. All authors have critically reviewed the manuscript and have approved the final version for publication. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AAVsAdeno-associated viruses
ABEsAdenosine base editors
ADAR2Adenosine deaminase domain of adenosine deaminase acting on the RNA-2
AdsAdenoviral vectors
CRISPR-Cas9Clustered Regularly Interspaced Short Palindromic Repeats CRISPR-Cas9 nuclease
CBEsCytosine base editors
CTLCytotoxic T lymphocyte
dCas9Dead Cas9
DDRDNA damage response
DSBsDouble-strand breaks
FAFanconi anemia
HbFFetal hemoglobin
HDAdHelper-dependent adenovirus
HSCHematopoietic stem cell
HSPCsHematopoietic stem and progenitor cells
HPFHHereditary persistence of fetal hemoglobin
HDRHomologous directed repair
HIVHuman immunodeficiency virus
IAVInfluenza A virus
ITRsInverted terminal repeats
LbLLayer-by-layer
LVsLentiviral vectors lvs
LCMVLymphocytic choriomeningitis virus
MHCMajor histocompatibility complex
MMEJMicrohomology-mediated end joining
NabNeutralizing antibodies
NHEJNon-homologous end joining
PNAsPeptide nucleic acids
PLGAPoly-lactic-co-glycolic acid
pegRNAPrime editing guide RNA
PEPrime editors
PAMProtospacer adjacent motif
SARS-CoV-2Severe acute respiratory syndrome coronavirus 2
ssODNSingle-stranded oligodeoxynucleotide
TALENsTranscription activator-like effector nucleases
VOCsVaso-occlusive crises
VSVVesicular stomatitis virus
VSV-GVesicular stomatitis virus glycoprotein
WASWiskott-Aldrich syndrome
XLMTMX-linked myotubular myopathy and Duchenne’s muscular dystrophy
X-SCIDX-linked Severe Combined Immunodeficiency
ZFNsZinc-finger nucleases

References

  1. Branzei, D.; Foiani, M. Regulation of DNA Repair throughout the Cell Cycle. Nat. Rev. Mol. Cell Biol. 2008, 9, 297–308. [Google Scholar] [CrossRef] [PubMed]
  2. Saleh-Gohari, N.; Helleday, T. Conservative Homologous Recombination Preferentially Repairs DNA Double-Strand Breaks in the S Phase of the Cell Cycle in Human Cells. Nucleic Acids Res. 2004, 32, 3683–3688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Ho, B.X.; Loh, S.J.H.; Chan, W.K.; Soh, B.S. In Vivo Genome Editing as a Therapeutic Approach. Int. J. Mol. Sci. 2018, 19, 2721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Dasgupta, I.; Flotte, T.R.; Keeler, A.M. CRISPR/Cas-Dependent and Nuclease-Free In Vivo Therapeutic Gene Editing. Hum. Gene Ther. 2021, 32, 275–293. [Google Scholar] [CrossRef] [PubMed]
  5. Cassandri, M.; Smirnov, A.; Novelli, F.; Pitolli, C.; Agostini, M.; Malewicz, M.; Melino, G.; Raschellà, G. Zinc-Finger Proteins in Health and Disease. Cell Death Discov. 2017, 3, 17071. [Google Scholar] [CrossRef] [Green Version]
  6. Singh, J.K.; van Attikum, H. DNA Double-Strand Break Repair: Putting Zinc Fingers on the Sore Spot. Semin. Cell Dev. Biol. 2021, 113, 65–74. [Google Scholar] [CrossRef]
  7. Hoban, M.D.; Cost, G.J.; Mendel, M.C.; Romero, Z.; Kaufman, M.L.; Joglekar, A.V.; Ho, M.; Lumaquin, D.; Gray, D.; Lill, G.R.; et al. Correction of the Sickle Cell Disease Mutation in Human Hematopoietic Stem/Progenitor Cells. Blood 2015, 125, 2597–2604. [Google Scholar] [CrossRef]
  8. Smith, E.C.; Luc, S.; Croney, D.M.; Woodworth, M.B.; Greig, L.C.; Fujiwara, Y.; Nguyen, M.; Sher, F.; Macklis, J.D.; Bauer, D.E.; et al. Strict in Vivo Specificity of the Bcl11a Erythroid Enhancer. Blood 2016, 128, 2338–2342. [Google Scholar] [CrossRef]
  9. Psatha, N.; Reik, A.; Phelps, S.; Zhou, Y.; Dalas, D.; Yannaki, E.; Levasseur, D.N.; Urnov, F.D.; Holmes, M.C.; Papayannopoulou, T. Disruption of the BCL11A Erythroid Enhancer Reactivates Fetal Hemoglobin in Erythroid Cells of Patients with β-Thalassemia Major. Mol. Ther.-Methods Clin. Dev. 2018, 10, 313–326. [Google Scholar] [CrossRef] [Green Version]
  10. Chang, K.H.; Smith, S.E.; Sullivan, T.; Chen, K.; Zhou, Q.; West, J.A.; Liu, M.; Liu, Y.; Vieira, B.F.; Sun, C.; et al. Long-Term Engraftment and Fetal Globin Induction upon BCL11A Gene Editing in Bone-Marrow-Derived CD34+ Hematopoietic Stem and Progenitor Cells. Mol. Ther.-Methods Clin. Dev. 2017, 4, 137–148. [Google Scholar] [CrossRef]
  11. Vierstra, J.; Reik, A.; Chang, K.H.; Stehling-Sun, S.; Zhou, Y.; Hinkley, S.J.; Paschon, D.E.; Zhang, L.; Psatha, N.; Bendana, Y.R.; et al. Functional Footprinting of Regulatory DNA. Nat. Methods 2015, 12, 927–930. [Google Scholar] [CrossRef] [PubMed]
  12. Sangamo, Sanofi Show Positive Early Data for SCD Gene-Edited Cell Therapy. Available online: https://www.genengnews.com/news/sangamo-sanofi-show-positive-early-data-for-scd-gene-edited-cell-therapy/ (accessed on 21 March 2021).
  13. Fox, T.A.; Booth, C. Gene Therapy for Primary Immunodeficiencies. Br. J. Haematol. 2021, 193, 1044–1059. [Google Scholar] [CrossRef] [PubMed]
  14. Genovese, P.; Schiroli, G.; Escobar, G.; Di Tomaso, T.; Firrito, C.; Calabria, A.; Moi, D.; Mazzieri, R.; Bonini, C.; Holmes, M.C.; et al. Targeted Genome Editing in Human Repopulating Haematopoietic Stem Cells. Nature 2014, 510, 235–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Schiroli, G.; Ferrari, S.; Conway, A.; Jacob, A.; Capo, V.; Albano, L.; Plati, T.; Castiello, M.C.; Sanvito, F.; Gennery, A.R.; et al. Preclinical Modeling Highlights the Therapeutic Potential of Hematopoietic Stem Cell Gene Editing for Correction of SCID-X1. Sci. Transl. Med. 2017, 9, eaan0820. [Google Scholar] [CrossRef]
  16. Pavel-Dinu, M.; Wiebking, V.; Dejene, B.T.; Srifa, W.; Mantri, S.; Nicolas, C.E.; Lee, C.; Bao, G.; Kildebeck, E.J.; Punjya, N.; et al. Gene Correction for SCID-X1 in Long-Term Hematopoietic Stem Cells. Nat. Commun. 2019, 10, 1634. [Google Scholar] [CrossRef] [Green Version]
  17. Candotti, F. Gene Therapy for Wiskott-Aldrich Syndrome: Here to Stay. Lancet Haematol. 2019, 6, e230–e231. [Google Scholar] [CrossRef] [Green Version]
  18. Laskowski, T.J.; Van Caeneghem, Y.; Pourebrahim, R.; Ma, C.; Ni, Z.; Garate, Z.; Crane, A.M.; Li, X.S.; Liao, W.; Gonzalez-Garay, M.; et al. Gene Correction of IPSCs from a Wiskott-Aldrich Syndrome Patient Normalizes the Lymphoid Developmental and Functional Defects. Stem Cell Rep. 2016, 7, 139–148. [Google Scholar] [CrossRef] [Green Version]
  19. Peterson, C.W.; Wang, J.; Norman, K.K.; Norgaard, Z.K.; Humbert, O.; Tse, C.K.; Yan, J.J.; Trimble, R.G.; Shivak, D.A.; Rebar, E.J.; et al. Long-Term Multilineage Engraftment of Autologous Genome-Edited Hematopoietic Stem Cells in Nonhuman Primates. Blood 2016, 127, 2416–2426. [Google Scholar] [CrossRef] [Green Version]
  20. 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] [Green Version]
  21. Becker, S.; Boch, J. TALE and TALEN Genome Editing Technologies. Gene Genome Ed. 2021, 2, 100007. [Google Scholar] [CrossRef]
  22. Bhardwaj, A.; Nain, V. TALENs—an Indispensable Tool in the Era of CRISPR: A Mini Review. J. Genet. Eng. Biotechnol. 2021, 19, 125. [Google Scholar] [CrossRef] [PubMed]
  23. Li, H.; Yang, Y.; Hong, W.; Huang, M.; Wu, M.; Zhao, X. Applications of Genome Editing Technology in the Targeted Therapy of Human Diseases: Mechanisms, Advances and Prospects. Signal Transduct. Target. Ther. 2020, 5, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Lux, C.T.; 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. 2019, 12, 175–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Humbert, O.; Peterson, C.W.; Norgaard, Z.K.; Radtke, S.; Kiem, H.P. A Nonhuman Primate Transplantation Model to Evaluate Hematopoietic Stem Cell Gene Editing Strategies for β-Hemoglobinopathies. Mol. Ther.-Methods Clin. Dev. 2018, 8, 75–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Patsali, P.; Turchiano, G.; Papasavva, P.; Romito, M.; Loucari, C.C.; Stephanou, C.; Christou, S.; Sitarou, M.; Mussolino, C.; Cornu, T.I.; et al. Correction of IVS I-110(G>A) β-Thalassemia by CRISPR/Cas- And TALEN-Mediated Disruption of Aberrant Regulatory Elements in Human Hematopoietic Stem and Progenitor Cells. Haematologica 2019, 104, E497–E501. [Google Scholar] [CrossRef] [Green Version]
  27. Xu, P.; Tong, Y.; Liu, X.Z.; Wang, T.T.; Cheng, L.; Wang, B.Y.; Lv, X.; Huang, Y.; Liu, D.P. Both TALENs and CRISPR/Cas9 Directly Target the HBB IVS2-654 (C > T) Mutation in β-Thalassemiaderived IPSCs. Sci. Rep. 2015, 5, srep12065. [Google Scholar] [CrossRef] [Green Version]
  28. Fang, Y.; Cheng, Y.; Lu, D.; Gong, X.; Yang, G.; Gong, Z.; Zhu, Y.; Sang, X.; Fan, S.; Zhang, J.; et al. Treatment of Β654-Thalassaemia by TALENs in a Mouse Model. Cell Prolif. 2018, 51, e12491. [Google Scholar] [CrossRef] [Green Version]
  29. Menon, T.; Firth, A.L.; Scripture-Adams, D.D.; Galic, Z.; Qualls, S.J.; Gilmore, W.B.; Ke, E.; Singer, O.; Anderson, L.S.; Bornzin, A.R.; et al. Lymphoid Regeneration from Gene-Corrected SCID-X1 Subject-Derived IPSCs. Cell Stem Cell 2015, 16, 367–372. [Google Scholar] [CrossRef] [Green Version]
  30. Cellectis Presents Initial Preclinical Data on Two Novel Gene Therapies for Patients with RAG1 Severe Combined Immunodeficiency (SCID) and Hyper IgE Syndrome at ESGCT 2021. Available online: https://cellectis.com/en/press/cellectis-presents-initial-preclinical-data-on-two-novel-gene-therapies-for-patients-with-rag1-severe-combined-immunodeficiency-scid-and-hyper-ige-syndrome-at-esgct-2021 (accessed on 12 September 2022).
  31. Shi, B.; Li, J.; Shi, X.; Jia, W.; Wen, Y. TALEN-Mediated Knockout of CCR5 Confers Protection Against Infection of Human Immunodeficiency Virus. JAIDS J. Acquir. Immune Defic. Syndr. 2017, 74, 229–241. [Google Scholar] [CrossRef]
  32. Romito, M.; Juillerat, A.; Kok, Y.L.; Hildenbeutel, M.; Rhiel, M.; Andrieux, G.; Geiger, J.; Rudolph, C.; Mussolino, C.; Duclert, A.; et al. Preclinical Evaluation of a Novel TALEN Targeting CCR5 Confirms Efficacy and Safety in Conferring Resistance to HIV-1 Infection. Biotechnol. J. 2021, 16, e2000023. [Google Scholar] [CrossRef]
  33. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A Programmable Dual-RNA-Guided DNA Endonuclease in Adaptive Bacterial Immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef] [PubMed]
  34. Doudna, J.A.; Charpentier, E. The New Frontier of Genome Engineering with CRISPR-Cas9. Science 2014, 346, 1258096. [Google Scholar] [CrossRef] [PubMed]
  35. Jiang, F.; Doudna, J.A. CRISPR–Cas9 Structures and Mechanisms. Annu. Rev. Biophys. 2017, 46, 505–531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Traxler, E.A.; Yao, Y.; Wang, Y.D.; Woodard, K.J.; Kurita, R.; Nakamura, Y.; Hughes, J.R.; Hardison, R.C.; Blobel, G.A.; Li, C.; et al. A Genome-Editing Strategy to Treat β-Hemoglobinopathies That Recapitulates a Mutation Associated with a Benign Genetic Condition. Nat. Med. 2016, 22, 987–990. [Google Scholar] [CrossRef]
  37. Métais, J.Y.; Doerfler, P.A.; Mayuranathan, T.; Bauer, D.E.; Fowler, S.C.; Hsieh, M.M.; Katta, V.; Keriwala, S.; Lazzarotto, C.R.; Luk, K.; et al. Genome Editing of HBG1 and HBG2 to Induce Fetal Hemoglobin. Blood Adv. 2019, 3, 3379–3392. [Google Scholar] [CrossRef] [Green Version]
  38. Antoniani, C.; Meneghini, V.; Lattanzi, A.; Felix, T.; Romano, O.; Magrin, E.; Weber, L.; Pavani, G.; El Hoss, S.; Kurita, R.; et al. Induction of Fetal Hemoglobin Synthesis by CRISPR/Cas9-Mediated Editing of the Human b-Globin Locus. Blood 2018, 131, 1960–1973. [Google Scholar] [CrossRef] [Green Version]
  39. Bauer, D.E.; Kamran, S.C.; Orkin, S.H. Reawakening Fetal Hemoglobin: Prospects for New Therapies for the β-Globin Disorders. Blood 2012, 120, 2945–2953. [Google Scholar] [CrossRef] [Green Version]
  40. Liu, N.; Xu, S.; Yao, Q.; Zhu, Q.; Kai, Y.; Hsu, J.Y.; Sakon, P.; Pinello, L.; Yuan, G.C.; Bauer, D.E.; et al. Transcription Factor Competition at the γ-Globin Promoters Controls Hemoglobin Switching. Nat. Genet. 2021, 53, 511–520. [Google Scholar] [CrossRef]
  41. Masuda, T.; Wang, X.; Maeda, M.; Canver, M.C.; Sher, F.; Funnell, A.P.W.; Fisher, C.; Suciu, M.; Martyn, G.E.; Norton, L.J.; et al. Gene Regulation: Transcription Factors LRF and BCL11A Independently Repress Expression of Fetal Hemoglobin. Science 2016, 351, 285–289. [Google Scholar] [CrossRef] [Green Version]
  42. Lamsfus-Calle, A.; Daniel-Moreno, A.; Antony, J.S.; Epting, T.; Heumos, L.; Baskaran, P.; Admard, J.; Casadei, N.; Latifi, N.; Siegmund, D.M.; et al. Comparative Targeting Analysis of KLF1, BCL11A, and HBG1/2 in CD34+ HSPCs by CRISPR/Cas9 for the Induction of Fetal Hemoglobin. Sci. Rep. 2020, 10, 10133. [Google Scholar] [CrossRef]
  43. Demirci, S.; Zeng, J.; Wu, Y.; Uchida, N.; Shen, A.H.; Pellin, D.; Gamer, J.; Yapundich, M.; Drysdale, C.; Bonanno, J.; et al. BCL11A Enhancer–Edited Hematopoietic Stem Cells Persist in Rhesus Monkeys without Toxicity. J. Clin. Invest. 2020, 130, 6677–6687. [Google Scholar] [CrossRef] [PubMed]
  44. Psatha, N.; Georgakopoulou, A.; Li, C.; Nandakumar, V.; Georgolopoulos, G.; Acosta, R.; Paschoudi, K.; Nelson, J.; Chee, D.; Athanasiadou, A.; et al. Enhanced HbF Reactivation by Multiplex Mutagenesis of Thalassemic CD34+ Cells in Vitro and in Vivo. Blood 2021, 138, 1540–1553. [Google Scholar] [CrossRef] [PubMed]
  45. Wu, Y.; Zeng, J.; Roscoe, B.P.; Liu, P.; Yao, Q.; Lazzarotto, C.R.; Clement, K.; Cole, M.A.; Luk, K.; Baricordi, C.; et al. Highly Efficient Therapeutic Gene Editing of Human Hematopoietic Stem Cells. Nat. Med. 2019, 25, 776–783. [Google Scholar] [CrossRef] [PubMed]
  46. Locatelli, F.; Frangoul, H.; Corbacioglu, S.; de la Fuente, J.; Wall, D.; Capellini, M.D.; de Montalembert, M.; Kattamis, A.; Lobitz, S.; Rondelli, D.; et al. Efficacy and Safety of a Single Dose of Ctx001 For Transfusion-Dependent Βeta-Thalassemia And Severe Sickle Cell Disease. In Proceedings of the Conference of European Hematology Association, Vienna, Austria, 9–17 June 2022. [Google Scholar]
  47. ClinicalTrials.Gov. Available online: https://clinicaltrials.gov/ct2/home (accessed on 20 June 2022).
  48. Román-Rodríguez, F.J.; Ugalde, L.; Álvarez, L.; Díez, B.; Ramírez, M.J.; Risueño, C.; Cortón, M.; Bogliolo, M.; Bernal, S.; March, F.; et al. NHEJ-Mediated Repair of CRISPR-Cas9-Induced DNA Breaks Efficiently Corrects Mutations in HSPCs from Patients with Fanconi Anemia. Cell Stem Cell 2019, 25, 607–621.e7. [Google Scholar] [CrossRef] [Green Version]
  49. Rai, R.; Romito, M.; Rivers, E.; Turchiano, G.; Blattner, G.; Vetharoy, W.; Ladon, D.; Andrieux, G.; Zhang, F.; Zinicola, M.; et al. Targeted Gene Correction of Human Hematopoietic Stem Cells for the Treatment of Wiskott-Aldrich Syndrome. Nat. Commun. 2020, 11, 4034. [Google Scholar] [CrossRef]
  50. Hou, P.; Chen, S.; Wang, S.; Yu, X.; Chen, Y.; Jiang, M.; Zhuang, K.; Ho, W.; Hou, W.; Huang, J.; et al. Genome Editing of CXCR4 by CRISPR/Cas9 Confers Cells Resistant to HIV-1 Infection. Sci. Rep. 2015, 5, 15577. [Google Scholar] [CrossRef] [Green Version]
  51. Yu, S.; Yao, Y.; Xiao, H.; Li, J.; Liu, Q.; Yang, Y.; Adah, D.; Lu, J.; Zhao, S.; Qin, L.; et al. Simultaneous Knockout of CXCR4 and CCR5 Genes in CD4+ T Cells via CRISPR/Cas9 Confers Resistance to Both X4- and R5-Tropic Human Immunodeficiency Virus Type 1 Infection. Hum. Gene Ther. 2018, 29, 51–67. [Google Scholar] [CrossRef]
  52. Xu, L.; Wang, J.; Liu, Y.; Xie, L.; Su, B.; Mou, D.; Wang, L.; Liu, T.; Wang, X.; Zhang, B.; et al. CRISPR-Edited Stem Cells in a Patient with HIV and Acute Lymphocytic Leukemia. N. Engl. J. Med. 2019, 381, 1240–1247. [Google Scholar] [CrossRef]
  53. Haapaniemi, E.; Botla, S.; Persson, J.; Schmierer, B.; Taipale, J. CRISPR-Cas9 Genome Editing Induces a P53-Mediated DNA Damage Response. Nat. Med. 2018, 24, 927–930. [Google Scholar] [CrossRef] [Green Version]
  54. Mirgayazova, R.; Khadiullina, R.; Chasov, V.; Mingaleeva, R.; Miftakhova, R.; Rizvanov, A.; Bulatov, E. Therapeutic Editing of the TP53 Gene: Is Crispr/CAS9 an Option? Genes 2020, 11, 704. [Google Scholar] [CrossRef]
  55. Schiroli, G.; Conti, A.; Ferrari, S.; della Volpe, L.; Jacob, A.; Albano, L.; Beretta, S.; Calabria, A.; Vavassori, V.; Gasparini, P.; et al. Precise Gene Editing Preserves Hematopoietic Stem Cell Function Following Transient P53-Mediated DNA Damage Response. Cell Stem Cell 2019, 24, 551–565.e8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Leibowitz, M.L.; Papathanasiou, S.; Doerfler, P.A.; Blaine, L.J.; Sun, L.; Yao, Y.; Zhang, C.Z.; Weiss, M.J.; Pellman, D. Chromothripsis as an On-Target Consequence of CRISPR–Cas9 Genome Editing. Nat. Genet. 2021, 53, 895–905. [Google Scholar] [CrossRef] [PubMed]
  57. Cullot, G.; Boutin, J.; Toutain, J.; Prat, F.; Pennamen, P.; Rooryck, C.; Teichmann, M.; Rousseau, E.; Lamrissi-Garcia, I.; Guyonnet-Duperat, V.; et al. CRISPR-Cas9 Genome Editing Induces Megabase-Scale Chromosomal Truncations. Nat. Commun. 2019, 10, 1136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Zuccaro, M.V.; Xu, J.; Mitchell, C.; Marin, D.; Zimmerman, R.; Rana, B.; Weinstein, E.; King, R.T.; Palmerola, K.L.; Smith, M.E.; et al. Allele-Specific Chromosome Removal after Cas9 Cleavage in Human Embryos. Cell 2020, 183, 1650–1664.e15. [Google Scholar] [CrossRef] [PubMed]
  59. Stephens, P.J.; Greenman, C.D.; Fu, B.; Yang, F.; Bignell, G.R.; Mudie, L.J.; Pleasance, E.D.; Lau, K.W.; Beare, D.; Stebbings, L.A.; et al. Massive Genomic Rearrangement Acquired in a Single Catastrophic Event during Cancer Development. Cell 2011, 144, 27–40. [Google Scholar] [CrossRef] [PubMed]
  60. Kim, D.; Luk, K.; Wolfe, S.A.; Kim, J.S. Evaluating and Enhancing Target Specificity of Gene-Editing Nucleases and Deaminases. Annu. Rev. Biochem. 2019, 88, 191–220. [Google Scholar] [CrossRef]
  61. Yang, Y.; Xu, J.; Ge, S.; Lai, L. CRISPR/Cas: Advances, Limitations, and Applications for Precision Cancer Research. Front. Med. 2021, 8, 649896. [Google Scholar] [CrossRef]
  62. 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]
  63. Porto, E.M.; Komor, A.C.; Slaymaker, I.M.; Yeo, G.W. Base Editing: Advances and Therapeutic Opportunities. Nat. Rev. Drug Discov. 2020, 19, 839–859. [Google Scholar] [CrossRef]
  64. Gaudelli, N.M.; Komor, A.C.; Rees, H.A.; Packer, M.S.; Badran, A.H.; Bryson, D.I.; Liu, D.R. Programmable Base Editing of T to G C in Genomic DNA without DNA Cleavage. Nature 2017, 551, 464–471. [Google Scholar] [CrossRef]
  65. 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] [Green Version]
  66. Nishimasu, H.; Shi, X.; Ishiguro, S.; Gao, L.; Hirano, S.; Okazaki, S.; Noda, T.; Abudayyeh, O.O.; Gootenberg, J.S.; Mori, H.; et al. Engineered CRISPR-Cas9 Nuclease with Expanded Targeting Space. Science 2018, 9, 1259–1262. [Google Scholar] [CrossRef] [PubMed]
  67. Kantor, A.; Mcclements, M.E. CRISPR-Cas9 DNA Base-Editing and Prime-Editing. Int. J. Mol. Sci. 2020, 21, 6240. [Google Scholar] [CrossRef] [PubMed]
  68. Gaudelli, N.M.; Lam, D.K.; Rees, H.A.; Solá-Esteves, N.M.; Barrera, L.A.; Born, D.A.; Edwards, A.; Gehrke, J.M.; Lee, S.J.; Liquori, A.J.; et al. Directed Evolution of Adenine Base Editors with Increased Activity and Therapeutic Application. Nat. Biotechnol. 2020, 38, 892–900. [Google Scholar] [CrossRef] [PubMed]
  69. Zeng, J.; Wu, Y.; Ren, C.; Bonanno, J.; Shen, A.H.; Shea, D.; Gehrke, J.M.; Clement, K.; Luk, K.; Yao, Q.; et al. Therapeutic Base Editing of Human Hematopoietic Stem Cells. Nat. Med. 2020, 26, 535–541. [Google Scholar] [CrossRef]
  70. Wang, L.; Li, L.; Ma, Y.; Hu, H.; Li, Q.; Yang, Y.; Liu, W.; Yin, S.; Li, W.; Fu, B.; et al. Reactivation of γ-Globin Expression through Cas9 or Base Editor to Treat β-Hemoglobinopathies. Cell Res. 2020, 30, 276–278. [Google Scholar] [CrossRef]
  71. Antoniou, P.; Hardouin, G.; Martinucci, P.; Frati, G.; Felix, T.; Chalumeau, A.; Fontana, L.; Martin, J.; Masson, C.; Brusson, M.; et al. Base-Editing-Mediated Dissection of a γ -Globin Cis -Regulatory Element for the Therapeutic Reactivation of Fetal Hemoglobin Expression. Nat. Commun. 2022, 13, 6618. [Google Scholar] [CrossRef]
  72. Newby, G.A.; Yen, J.S.; Woodard, K.J.; Mayuranathan, T.; Lazzarotto, C.R.; Li, Y.; Sheppard-Tillman, H.; Porter, S.N.; Yao, Y.; Mayberry, K.; et al. Base Editing of Haematopoietic Stem Cells Rescues Sickle Cell Disease in Mice. Nature 2021, 595, 295–302. [Google Scholar] [CrossRef]
  73. Li, C.; Georgakopoulou, A.; Mishra, A.; Gil, S.; Hawkins, R.D.; Yannaki, E.; Lieber, A. In Vivo HSPC Gene Therapy with Base Editors Allows for Efficient Reactivation of Fetal G-Globin in b-YAC Mice. Blood Adv. 2021, 5, 1122–1135. [Google Scholar] [CrossRef]
  74. Li, J.; Zhou, Z.; Sun, H.X.; Ouyang, W.; Dong, G.; Liu, T.; Ge, L.; Zhang, X.; Liu, C.; Gu, Y. Transcriptome Analyses of β-Thalassemia −28(A>G) Mutation Using Isogenic Cell Models Generated by CRISPR/Cas9 and Asymmetric Single-Stranded Oligodeoxynucleotides (AssODNs). Front. Genet. 2020, 11, 577053. [Google Scholar] [CrossRef]
  75. Gehrke, J.M.; Cervantes, O.; Clement, M.K.; Wu, Y.; Zeng, J.; Bauer, D.E.; Pinello, L.; Joung, J.K. An Apobec3a-Cas9 Base Editor with Minimized Bystander and off-Target Activities. Nat. Biotechnol. 2018, 36, 977. [Google Scholar] [CrossRef] [PubMed]
  76. Knipping, F.; Newby, G.A.; Eide, C.R.; McElroy, A.N.; Nielsen, S.C.; Smith, K.; Fang, Y.; Cornu, T.I.; Costa, C.; Gutierrez-Guerrero, A.; et al. Disruption of HIV-1 Co-Receptors CCR5 and CXCR4 in Primary Human T Cells and Hematopoietic Stem and Progenitor Cells Using Base Editing. Mol. Ther. 2022, 30, 130–144. [Google Scholar] [CrossRef] [PubMed]
  77. Siegner, S.M.; Clemens, A.; Ugalde, L.; Garcia-Garcia, L.; Bueren, J.A.; Rio, P.; Karasu, M.E.; Corn, J.E. Adenine Base Editing Is an Efficient Approach to Restore Function in FA Patient Cells without Double-Stranded DNA Breaks. bioRxiv 2022, 2022.04.22.489197. [Google Scholar] [CrossRef]
  78. McAuley, G.E.; Yiu, G.; Newby, G.A.; Kang, S.H.L.; Garibay, A.J.; Butler, J.A.; Christian, V.S.; Fitz-Gibbon, S.; Wong, R.L.; Everette, K.A.; et al. Base Editing of Hematopoietic Stem Cells Rescues T-Cell Development for CD3d Severe Combined Immunodeficiency. In Proceedings of the American Society of Gene and Cell Therapy, Annual Meeting, Washington, DC, USA, 16–19 May 2022. [Google Scholar]
  79. Schene, I.F.; Joore, I.P.; Oka, R.; Mokry, M.; van Vugt, A.H.M.; van Boxtel, R.; van der Doef, H.P.J.; van der Laan, L.J.W.; Verstegen, M.M.A.; van Hasselt, P.M.; et al. Prime Editing for Functional Repair in Patient-Derived Disease Models. Nat. Commun. 2020, 11, 5352. [Google Scholar] [CrossRef] [PubMed]
  80. Anzalone, A.V.; Koblan, L.W.; Liu, D.R. Genome Editing with CRISPR–Cas Nucleases, Base Editors, Transposases and Prime Editors. Nat. Biotechnol. 2020, 38, 824–844. [Google Scholar] [CrossRef] [PubMed]
  81. Chen, P.J.; Hussmann, J.A.; Yan, J.; Knipping, F.; Ravisankar, P.; Chen, P.-F.; Chen, C.; Nelson, J.W.; Newby, G.A.; Sahin, M.; et al. Enhanced Prime Editing Systems by Manipulating Cellular Determinants of Editing Outcomes. Cell 2021, 184, 5635–5652.e29. [Google Scholar] [CrossRef]
  82. Bird, A. Perceptions of Epigenetics. Nature 2007, 447, 396–398. [Google Scholar] [CrossRef]
  83. Ferrand, J.; Plessier, A.; Polo, S.E. Control of the Chromatin Response to DNA Damage: Histone Proteins Pull the Strings. Semin. Cell Dev. Biol. 2021, 113, 75–87. [Google Scholar] [CrossRef]
  84. Rasmussen, K.D.; Helin, K. Role of TET Enzymes in DNA Methylation, Development, and Cancer. Genes Dev. 2016, 30, 733–750. [Google Scholar] [CrossRef] [Green Version]
  85. Milazzo, G.; Mercatelli, D.; Di Muzio, G.; Triboli, L.; De Rosa, P.; Perini, G.; Giorgi, F.M. Histone Deacetylases (HDACs): Evolution, Specificity, Role in Transcriptional Complexes, and Pharmacological Actionability. Genes 2020, 11, 556. [Google Scholar] [CrossRef]
  86. Taylor, E.L.; Westendorf, J.J. Histone Mutations and Bone Cancers. Adv. Exp. Med. Biol. 2021, 1283, 53–62. [Google Scholar] [PubMed]
  87. Syding, L.A.; Nickl, P.; Kasparek, P.; Sedlacek, R. CRISPR/Cas9 Epigenome Editing Potential for Rare Imprinting Diseases: A Review. Cells 2020, 9, 993. [Google Scholar] [CrossRef] [PubMed]
  88. Rivenbark, A.G.; Stolzenburg, S.; Beltran, A.S.; Yuan, X.; Rots, M.G.; Strahl, B.D.; Blancafort, P. Epigenetic Reprogramming of Cancer Cells via Targeted DNA Methylation. Epigenetics. 2012, 7, 350–360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Bernstein, D.L.; Le Lay, J.E.; Ruano, E.G.; Kaestner, K.H. TALE-Mediated Epigenetic Suppression of CDKN2A Increases Replication in Human Fibroblasts. J. Clin. Invest. 2015, 125, 1998–2006. [Google Scholar] [CrossRef]
  90. Huang, Y.-H.; Su, J.; Lei, Y.; Brunetti, L.; Gundry, M.C.; Zhang, X.; Jeong, M.; Li, W.; Goodell, M.A. DNA Epigenome Editing Using CRISPR-Cas SunTag-Directed DNMT3A. Genome Biol. 2017, 18, 176. [Google Scholar] [CrossRef] [Green Version]
  91. Xu, X.; Tao, Y.; Gao, X.; Zhang, L.; Li, X.; Zou, W.; Ruan, K.; Wang, F.; Xu, G.; Hu, R. A CRISPR-Based Approach for Targeted DNA Demethylation. Cell Discov. 2016, 2, 16009. [Google Scholar] [CrossRef] [Green Version]
  92. Amabile, A.; Migliara, A.; Capasso, P.; Biffi, M.; Cittaro, D.; Naldini, L.; Lombardo, A. Inheritable Silencing of Endogenous Genes by Hit-and-Run Targeted Epigenetic Editing. Cell 2016, 167, 219–232.e14. [Google Scholar] [CrossRef] [Green Version]
  93. 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. In Vivo Target Gene Activation via CRISPR/Cas9-Mediated Trans-Epigenetic Modulation. Cell 2017, 171, 1495–1507.e15. [Google Scholar] [CrossRef] [Green Version]
  94. Moreno, A.M.; Fu, X.; Zhu, J.; Katrekar, D.; Shih, Y.R.V.; Marlett, J.; Cabotaje, J.; Tat, J.; Naughton, J.; Lisowski, L.; et al. In Situ Gene Therapy via AAV-CRISPR-Cas9-Mediated Targeted Gene Regulation. Mol. Ther. 2018, 26, 1818–1827. [Google Scholar] [CrossRef] [Green Version]
  95. Nuñez, J.K.; Chen, J.; Pommier, G.C.; Cogan, J.Z.; Replogle, J.M.; Adriaens, C.; Ramadoss, G.N.; Shi, Q.; Hung, K.L.; Samelson, A.J.; et al. Genome-Wide Programmable Transcriptional Memory by CRISPR-Based Epigenome Editing. Cell 2021, 184, 2503–2519.e17. [Google Scholar] [CrossRef]
  96. Xu, C.; Zhou, Y.; Xiao, Q.; He, B.; Geng, G.; Wang, Z.; Cao, B.; Dong, X.; Bai, W.; Wang, Y.; et al. Programmable RNA Editing with Compact CRISPR–Cas13 Systems from Uncultivated Microbes. Nat. Methods 2021, 18, 499–506. [Google Scholar] [CrossRef] [PubMed]
  97. Abudayyeh, O.O.; Gootenberg, J.S.; Konermann, S.; Joung, J.; Slaymaker, I.M.; Cox, D.B.T.; Shmakov, S.; Makarova, K.S.; Semenova, E.; Minakhin, L.; et al. C2c2 Is a Single-Component Programmable RNA-Guided RNA-Targeting CRISPR Effector. Science 2016, 353, aaf5573. [Google Scholar] [CrossRef] [Green Version]
  98. 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] [Green Version]
  99. 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] [PubMed] [Green Version]
  100. Tang, T.; Han, Y.; Wang, Y.; Huang, H.; Qian, P. Programmable System of Cas13-Mediated RNA Modification and Its Biological and Biomedical Applications. Front. Cell Dev. Biol. 2021, 9, 677587. [Google Scholar] [CrossRef]
  101. Fukuda, M.; Umeno, H.; Nose, K.; Nishitarumizu, A.; Noguchi, R.; Nakagawa, H. Construction of a Guide-RNA for Site-Directed RNA Mutagenesis Utilising Intracellular A-To-I RNA Editing. Sci. Rep. 2017, 7, 41478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Freije, C.A.; Myhrvold, C.; Boehm, C.K.; Lin, A.E.; Welch, N.L.; Carter, A.; Metsky, H.C.; Luo, C.Y.; Abudayyeh, O.O.; Gootenberg, J.S.; et al. Programmable Inhibition and Detection of RNA Viruses Using Cas13. Mol. Cell 2019, 76, 826–837.e11. [Google Scholar] [CrossRef] [Green Version]
  103. Abbott, T.R.; Dhamdhere, G.; Liu, Y.; Lin, X.; Goudy, L.; Zeng, L.; Chemparathy, A.; Chmura, S.; Heaton, N.S.; Debs, R.; et al. Development of CRISPR as an Antiviral Strategy to Combat SARS-CoV-2 and Influenza. Cell 2020, 181, 865–876.e12. [Google Scholar] [CrossRef]
  104. Yin, L.; Zhao, F.; Sun, H.; Wang, Z.; Huang, Y.; Zhu, W.; Xu, F.; Mei, S.; Liu, X.; Zhang, D.; et al. CRISPR-Cas13a Inhibits HIV-1 Infection. Mol. Ther.-Nucleic Acids 2020, 21, 147–155. [Google Scholar] [CrossRef]
  105. Ide, L.M.; Gangadharan, B.; Chiang, K.Y.; Doering, C.B.; Spencer, H.T. Hematopoietic Stem-Cell Gene Therapy of Hemophilia A Incorporating a Porcine Factor VIII Transgene and Nonmyeloablative Conditioning Regimens. Blood 2007, 110, 2855–2863. [Google Scholar] [CrossRef]
  106. Naldini, L. Lentiviruses as Gene Transfer Agents for Delivery to Non-Dividing Cells. Curr. Opin. Biotechnol. 1998, 9, 457–463. [Google Scholar] [CrossRef] [PubMed]
  107. Schröder, A.R.; Shinn, P.; Chen, H.; Berry, C.; Ecker, J.R.; Bushman, F. HIV-1 Integration in the Human Genome Favors Active Genes and Local Hotspots. Cell 2002, 110, 521–529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Aiuti, A.; Biasco, L.; Scaramuzza, S.; Ferrua, F.; Cicalese, M.P.; Baricordi, C.; Dionisio, F.; Calabria, A.; Giannelli, S.; Castiello, M.C.; et al. Lentiviral Hematopoietic Stem Cell Gene Therapy in Patients with Wiskott-Aldrich Syndrome. Science 2013, 341, 1233151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Ferrua, F.; Cicalese, M.P.; Galimberti, S.; Giannelli, S.; Dionisio, F.; Barzaghi, F.; Migliavacca, M.; Bernardo, M.E.; Calbi, V.; Assanelli, A.A.; et al. Lentiviral Haemopoietic Stem/Progenitor Cell Gene Therapy for Treatment of Wiskott-Aldrich Syndrome: Interim Results of a Non-Randomised, Open-Label, Phase 1/2 Clinical Study. Lancet Haematol. 2019, 6, e239–e253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Mamcarz, E.; Therapy, C.; Zhou, S.; Lockey, T.; Production, T.; Abdelsamed, H.; Cross, S.J.; Kang, G.; Condori, J.; Dowdy, J.; et al. Lentiviral Gene Therapy Combined with Low-Dose Busulfan in Infants with SCID-X1. N. Engl. J. Med. 2019, 380, 1525–1534. [Google Scholar] [CrossRef] [PubMed]
  111. Kohn, D.B.; Booth, C.; Shaw, K.L.; Xu-Bayford, J.; Garabedian, E.; Trevisan, V.; Carbonaro-Sarracino, D.A.; Soni, K.; Terrazas, D.; Snell, K.; et al. Autologous Ex Vivo Lentiviral Gene Therapy for Adenosine Deaminase Deficiency. N. Engl. J. Med. 2021, 384, 2002–2013. [Google Scholar] [CrossRef]
  112. Thompson, A.A.; Walters, M.C.; Kwiatkowski, J.; Rasko, J.E.J.; Ribeil, J.-A.; Hongeng, S.; Magrin, E.; Schiller, G.J.; Payen, E.; Semeraro, M.; et al. Gene Therapy in Patients with Transfusion-Dependent β-Thalassemia. N. Engl. J. Med. 2018, 378, 1479–1493. [Google Scholar] [CrossRef] [PubMed]
  113. Biffi, A.; Montini, E.; Lorioli, L.; Cesani, M.; Fumagalli, F.; Plati, T.; Baldoli, C.; Martino, S.; Calabria, A.; Canale, S.; et al. Lentiviral Hematopoietic Stem Cell Gene Therapy Benefits Metachromatic Leukodystrophy. Science 2013, 341, 1233158. [Google Scholar] [CrossRef] [Green Version]
  114. Cartier, N.; Hacein-Bey-Abina, S.; Bartholomae, C.C.; Bougnres, P.; Schmidt, M.; Von Kalle, C.; Fischer, A.; Cavazzana-Calvo, M.; Aubourg, P. Lentiviral Hematopoietic Cell Gene Therapy for X-Linked Adrenoleukodystrophy. Methods Enzymol. 2012, 507, 187–198. [Google Scholar] [CrossRef]
  115. Naldini, L.; Blömer, U.; Gallay, P.; Ory, D.; Mulligan, R.; Gage, F.H.; Verma, I.M.; Trono, D. In Vivo Gene Delivery and Stable Transduction of Nondividing Cells by a Lentiviral Vector. Science 1996, 272, 263–267. [Google Scholar] [CrossRef]
  116. Annoni, A.; Goudy, K.; Akbarpour, M.; Naldini, L.; Roncarolo, M.G. Immune Responses in Liver-Directed Lentiviral Gene Therapy. Transl. Res. 2013, 161, 230–240. [Google Scholar] [CrossRef] [PubMed]
  117. Milani, M.; Annoni, A.; Bartolaccini, S.; Biffi, M.; Russo, F.; Di Tomaso, T.; Raimondi, A.; Lengler, J.; Holmes, M.C.; Scheiflinger, F.; et al. Genome Editing for Scalable Production of Alloantigen-free Lentiviral Vectors for in Vivo Gene Therapy. EMBO Mol. Med. 2017, 9, 1558–1573. [Google Scholar] [CrossRef] [PubMed]
  118. DePolo, N.J.; Harkleroad, C.E.; Bodner, M.; Watt, A.T.; Anderson, C.G.; Greengard, J.S.; Murthy, K.K.; Dubensky, T.W.; Jolly, D.J. The Resistance of Retroviral Vectors Produced from Human Cells to Serum Inactivation In Vivo and In Vitro Is Primate Species Dependent. J. Virol. 1999, 73, 6708–6714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Girard-Gagnepain, A.; Amirache, F.; Costa, C.; Lévy, C.; Frecha, C.; Fusil, F.; Nègre, D.; Lavillette, D.; Cosset, F.L.; Verhoeyen, E. Baboon Envelope Pseudotyped LVs Outperform VSV-G-LVs for Gene Transfer into Early-Cytokine-Stimulated and Resting HSCs. Blood 2014, 124, 1221–1231. [Google Scholar] [CrossRef] [Green Version]
  120. Hwang, B.Y.; Schaffer, D.V. Engineering a Serum-Resistant and Thermostable Vesicular Stomatitis Virus G Glycoprotein for Pseudotyping Retroviral and Lentiviral Vectors. Gene Ther. 2013, 20, 807–815. [Google Scholar] [CrossRef] [Green Version]
  121. Rajawat, Y.S.; Humbert, O.; Cook, S.M.; Radtke, S.; Pande, D.; Enstrom, M.; Wohlfahrt, M.E.; Kiem, H.P. In Vivo Gene Therapy for Canine SCID-X1 Using Cocal-Pseudotyped Lentiviral Vector. Hum. Gene Ther. 2021, 32, 113–127. [Google Scholar] [CrossRef]
  122. Colamartino, A.B.L.; Lemieux, W.; Bifsha, P.; Nicoletti, S.; Chakravarti, N.; Sanz, J.; Roméro, H.; Selleri, S.; Béland, K.; Guiot, M.; et al. Efficient and Robust NK-Cell Transduction With Baboon Envelope Pseudotyped Lentivector. Front. Immunol. 2019, 10, 2873. [Google Scholar] [CrossRef] [Green Version]
  123. Gutierrez-Guerrero, A.; Cosset, F.L.; Verhoeyen, E. Lentiviral Vector Pseudotypes: Precious Tools to Improve Gene Modification of Hematopoietic Cells for Research and Gene Therapy. Viruses 2020, 12, 1016. [Google Scholar] [CrossRef]
  124. Burtner, C.R.; Beard, B.C.; Kennedy, D.R.; Wohlfahrt, M.E.; Adair, J.E.; Trobridge, G.D.; Scharenberg, A.M.; Torgerson, T.R.; Rawlings, D.J.; Felsburg, P.J.; et al. Intravenous Injection of a Foamy Virus Vector to Correct Canine SCID-X1. Blood 2014, 123, 3578–3584. [Google Scholar] [CrossRef] [Green Version]
  125. Lévy, C.; Amirache, F.; Girard-Gagnepain, A.; Frecha, C.; Roman-Rodríguez, F.J.; Bernadin, O.; Costa, C.; Nègre, D.; Gutierrez-Guerrero, A.; Vranckx, L.S.; et al. Measles Virus Envelope Pseudotyped Lentiviral Vectors Transduce Quiescent Human HSCs at an Efficiency without Precedent. Blood Adv. 2017, 1, 2088–2104. [Google Scholar] [CrossRef]
  126. Ortinski, P.I.; O’Donovan, B.; Dong, X.; Kantor, B. Integrase-Deficient Lentiviral Vector as an All-in-One Platform for Highly Efficient CRISPR/Cas9-Mediated Gene Editing. Mol. Ther.-Methods Clin. Dev. 2017, 5, 153–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Hu, J.; Schokrpur, S.; Archang, M.; Hermann, K.; Sharrow, A.C.; Khanna, P.; Novak, J.; Signoretti, S.; Bhatt, R.S.; Knudsen, B.S.; et al. A Non-Integrating Lentiviral Approach Overcomes Cas9-Induced Immune Rejection to Establish an Immunocompetent Metastatic Renal Cancer Model. Mol. Ther.-Methods Clin. Dev. 2018, 9, 203–210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Mock, U.; Riecken, K.; Berdien, B.; Qasim, W.; Chan, E.; Cathomen, T.; Fehse, B. Novel Lentiviral Vectors with Mutated Reverse Transcriptase for MRNA Delivery of TALE Nucleases. Sci. Rep. 2014, 4, 6409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Burnett JR, H.A. Alipogene Tiparvovec, an Adeno-Associated Virus Encoding the Ser(447)X Variant of the Human Lipoprotein Lipase Gene for the Treatment of Patients with Lipoprotein Lipase Deficiency. Curr. Opin. Mol. Ther. 2009, 11, 681–691. [Google Scholar] [PubMed]
  130. Russell, S.; Bennett, J.; Wellman, J.A.; Chung, D.C.; Yu, Z.F.; Tillman, A.; Wittes, J.; Pappas, J.; Elci, O.; McCague, S.; et al. Efficacy and Safety of Voretigene Neparvovec (AAV2-HRPE65v2) in Patients with RPE65-Mediated Inherited Retinal Dystrophy: A Randomised, Controlled, Open-Label, Phase 3 Trial. Lancet 2017, 390, 849–860. [Google Scholar] [CrossRef] [PubMed]
  131. Hoy, S.M. Onasemnogene Abeparvovec: First Global Approval. Drugs 2019, 79, 1255–1262. [Google Scholar] [CrossRef] [PubMed]
  132. Podsakoff, G.; Wong, K.K.; Chatterjee, S. Efficient Gene Transfer into Nondividing Cells by Adeno-Associated Virus-Based Vectors. J. Virol. 1994, 68, 5656–5666. [Google Scholar] [CrossRef] [Green Version]
  133. Shirley, J.L.; de Jong, Y.P.; Terhorst, C.; Herzog, R.W. Immune Responses to Viral Gene Therapy Vectors. Mol. Ther. 2020, 28, 709–722. [Google Scholar] [CrossRef]
  134. Nathwani, A.C.; Hanawa, H.; Vandergriff, J.; Kelly, P.; Vanin, E.F.; Nienhuis, A.W. Efficient Gene Transfer into Human Cord Blood CD34+ Cells and the CD34+CD38- Subset Using Highly Purified Recombinant Adeno-Associated Viral Vector Preparations That Are Free of Helper Virus and Wild-Type AAV. Gene Ther. 2000, 7, 183–195. [Google Scholar] [CrossRef] [Green Version]
  135. Santat, L.; Paz, H.; Wong, C.; Li, L.; Macer, J.; Forman, S.; Wong, K.K.; Chatterjee, S. Recombinant AAV2 Transduction of Primitive Human Hematopoietic Stem Cells Capable of Serial Engraftment in Immune-Deficient Mice. Proc. Natl. Acad. Sci. USA 2005, 102, 11053–11058. [Google Scholar] [CrossRef]
  136. Zhou, S.Z.; Cooper, S.; Kang, L.Y.; Ruggieri, L.; Heimfeld, S.; Srivastava, A.; Broxmeyer, H.E. Adeno-Associated Virus 2-Mediated High Efficiency Gene Transfer into Irm, Nature and Mature Subsets of Hematopoietic Progenitor Cells in Human Umbilical Cord Blood. J. Exp. Med. 1994, 179, 1867–1875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Ponnazhagan, S.; Mukherjee, P.; Wang, X.S.; Qing, K.; Kube, D.M.; Mah, C.; Kurpad, C.; Yoder, M.C.; Srour, E.F.; Srivastava, A. Adeno-Associated Virus Type 2-Mediated Transduction in Primary Human Bone Marrow-Derived CD34+ Hematopoietic Progenitor Cells: Donor Variation and Correlation of Transgene Expression with Cellular Differentiation. J. Virol. 1997, 71, 8262–8267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Hargrove, P.W.; Vanin, E.F.; Kurtzman, G.J.; Nienhuis, A.W. High-Level Globin Gene Expression Mediated by a Recombinant Adeno- Associated Virus Genome That Contains the 3* g Globin Gene Regulatory Element and Integrates as Tandem Copies in Erythroid Cells. Red Cells 1997, 89, 2167–2175. [Google Scholar] [CrossRef]
  139. Malik, P.; McQuiston, S.A.; Yu, X.J.; Pepper, K.A.; Krall, W.J.; Podsakoff, G.M.; Kurtzman, G.J.; Kohn, D.B. Recombinant Adeno-Associated Virus Mediates a High Level of Gene Transfer but Less Efficient Integration in the K562 Human Hematopoietic Cell Line. J. Virol. 1997, 71, 1776–1783. [Google Scholar] [CrossRef] [Green Version]
  140. Song, L.; Kauss, M.A.; Kopin, E.; Chandra, M.; Ul-hasan, T.; Miller, E.; Jayandharan, G.R.; Rivers, A.E.; Aslanidi, G.V.; Ling, C.; et al. Optimizing the Transduction Efficiency of Human Hematopoietic Stem Cells Using Capsid-Modified AAV6 Vectors in Vitro and in a Xenograft Mouse Model in Vivo. Cytotherapy 2014, 15, 986–998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Pattabhi, S.; Lotti, S.N.; Berger, M.P.; Singh, S.; Lux, C.T.; Jacoby, K.; Lee, C.; Negre, O.; Scharenberg, A.M.; Rawlings, D.J. In Vivo Outcome of Homology-Directed Repair at the HBB Gene in HSC Using Alternative Donor Template Delivery Methods. Mol. Ther.-Nucleic Acids 2019, 17, 277–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Wilkinson, A.C.; Dever, D.P.; Baik, R.; Camarena, J.; Hsu, I.; Charlesworth, C.T.; Morita, C.; Nakauchi, H.; Porteus, M.H. Cas9-AAV6 Gene Correction of β-Globin in Autologous HSCs Improves Sickle Cell Disease Erythropoiesis in Mice. Nat. Commun. 2021, 12, 686. [Google Scholar] [CrossRef] [PubMed]
  143. Romero, Z.; Lomova, A.; Said, S.; Miggelbrink, A.; Kuo, C.Y.; Campo-Fernandez, B.; Hoban, M.D.; Masiuk, K.E.; Clark, D.N.; Long, J.; et al. Editing the Sickle Cell Disease Mutation in Human Hematopoietic Stem Cells: Comparison of Endonucleases and Homologous Donor Templates. Mol. Ther. 2019, 27, 1389–1406. [Google Scholar] [CrossRef]
  144. Wang, J.; Exline, C.M.; Declercq, J.J.; Llewellyn, G.N.; Hayward, B.; Li, P.W.; Shivak, D.A.; Surosky, R.T.; Philip, D.; Holmes, M.C.; et al. Homology-Driven Genome Editing in Hematopoietic Stem and Progenitor Cells Using Zinc Finger Nuclease MRNA and AAV6 Donors. Nat. Biotechnol. 2015, 33, 1256–1263. [Google Scholar] [CrossRef] [Green Version]
  145. Martin, R.M.; Ikeda, K.; Cromer, M.K.; Uchida, N.; Nishimura, T.; Romano, R.; Tong, A.J.; Lemgart, V.T.; Camarena, J.; Pavel-Dinu, M.; et al. Highly Efficient and Marker-Free Genome Editing of Human Pluripotent Stem Cells by CRISPR-Cas9 RNP and AAV6 Donor-Mediated Homologous Recombination. Cell Stem Cell 2019, 24, 821–828.e5. [Google Scholar] [CrossRef]
  146. Ellis, B.L.; Hirsch, M.L.; Barker, J.C.; Connelly, J.P.; Steininger, R.J.; Porteus, M.H. A Survey of Ex Vivo/in Vitro Transduction Efficiency of Mammalian Primary Cells and Cell Lines with Nine Natural Adeno-Associated Virus (AAV1-9) and One Engineered Adeno-Associated Virus Serotype. Virol. J. 2013, 10, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. 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 Hemostasis in a Mouse Model of Hemophilia. Nature 2012, 475, 217–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. 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] [Green Version]
  149. 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] [PubMed] [Green Version]
  150. Krooss, S.A.; Dai, Z.; Schmidt, F.; Rovai, A.; Fakhiri, J.; Dhingra, A.; Yuan, Q.; Yang, T.; Balakrishnan, A.; Steinbrück, L.; et al. Ex Vivo/In Vivo Gene Editing in Hepatocytes Using “All-in-One” CRISPR-Adeno-Associated Virus Vectors with a Self-Linearizing Repair Template. iScience 2020, 23, 100764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Thakore, P.I.; Kwon, J.B.; Nelson, C.E.; Rouse, D.C.; Gemberling, M.P.; Oliver, M.L.; Gersbach, C.A. RNA-Guided Transcriptional Silencing in Vivo with S. Aureus CRISPR-Cas9 Repressors. Nat. Commun. 2018, 9, 1674. [Google Scholar] [CrossRef] [Green Version]
  152. Ran, F.A.; Cong, L.; Yan, W.X.; Scott, D.A.; Gootenberg, J.S.; Kriz, A.J.; Zetsche, B.; Shalem, O.; Wu, X.; Makarova, K.S.; et al. In Vivo Genome Editing Using Staphylococcus Aureus Cas9. Nature 2015, 520, 186–191. [Google Scholar] [CrossRef] [Green Version]
  153. Calcedo, R.; Morizono, H.; Wang, L.; McCarter, R.; He, J.; Jones, D.; Batshaw, M.L.; Wilson, J.M. Adeno-Associated Virus Antibody Profiles in Newborns, Children, and Adolescents. Clin. Vaccine Immunol. 2011, 18, 1586–1588. [Google Scholar] [CrossRef] [Green Version]
  154. Rapti, K.; Louis-Jeune, V.; Kohlbrenner, E.; Ishikawa, K.; Ladage, D.; Zolotukhin, S.; Hajjar, R.J.; Weber, T. Neutralizing Antibodies against AAV Serotypes 1, 2, 6, and 9 in Sera of Commonly Used Animal Models. Mol. Ther. 2012, 20, 73–83. [Google Scholar] [CrossRef] [Green Version]
  155. Manno, C.S.; Chew, A.J.; Hutchison, S.; Larson, P.J.; Herzog, R.W.; Arruda, V.R.; Tai, S.J.; Ragni, M.V.; Thompson, A.; Ozelo, M.; et al. AAV-Mediated Factor IX Gene Transfer to Skeletal Muscle in Patients with Severe Hemophilia B. Blood 2003, 101, 2963–2972. [Google Scholar] [CrossRef]
  156. Mingozzi, F.; Anguela, X.M.; Pavani, G.; Chen, Y.; Davidson, R.J.; Hui, D.J.; Yazicioglu, M.; Elkouby, L.; Hinderer, C.J.; Faella, A.; et al. Overcoming Pre-Existing Humoral Immunity to AAV Using Capsid Decoys. Sci. Transl. Med. 2013, 5, S45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Greenberg, B.H.; A Butler, J.; Felker, G.; Ponikowski, P.; A Voors, A.; Pogoda, J.M.; Provost, R.; Guerrero, J.L.; Hajjar, R.J.; Zsebo, K.M. Prevalence of AAV1 Neutralizing Antibodies and Consequences for a Clinical Trial of Gene Transfer for Advanced Heart Failure. Gene Ther. 2016, 23, 313–319. [Google Scholar] [CrossRef] [PubMed]
  158. Majowicz, A.; Nijmeijer, B.; Lampen, M.H.; Spronck, L.; de Haan, M.; Petry, H.; van Deventer, S.J.; Meyer, C.; Tangelder, M.; Ferreira, V. Therapeutic HFIX Activity Achieved after Single AAV5-HFIX Treatment in Hemophilia B Patients and NHPs with Pre-Existing Anti-AAV5 NABs. Mol. Ther.-Methods Clin. Dev. 2019, 14, 27–36. [Google Scholar] [CrossRef] [Green Version]
  159. Wilson, J.M.; Flotte, T.R. Moving Forward after Two Deaths in a Gene Therapy Trial of Myotubular Myopathy. Hum. Gene Ther. 2020, 31, 695–696. [Google Scholar] [CrossRef] [PubMed]
  160. Taylor, P. Astellas Reports Fourth Death in Halted Gene Therapy Trial. Available online: https://www.medscape.com/viewarticle/979152 (accessed on 12 September 2022).
  161. Burton, K.W. Two Deaths From Liver Failure Linked to Spinal Muscular Atrophy Drug. Available online: https://pharmaphorum.com/news/astellas-reports-fourth-death-in-halted-gene-therapy-trial/ (accessed on 12 September 2022).
  162. Rowe, W.P.; Huebner, R.J.; Gilmore, L.K.; Parrott, R.H.; Ward, T.G. Isolation of a Cytopathogenic Agent from Human Adenoids LTndergoing Spontaneous Degeneration in Tissue Culture. World Health Organ. Monogr. Ser. 1952, 64, 84. [Google Scholar]
  163. Rosenfeld, M.A.; Siegfried, W.; Yoshimura, K.; Yoneyama, K.; Fukayama, M.; Stier, L.E.; Pääkkö, P.K.; Gilardi, P.; Stratford-Perricaudet, L.D.; Perricaudet, M.; et al. Adenovirus-Mediated Transfer of a Recombinant A1-Antitrypsin Gene to the Lung Epithelium in Vivo. Science 1991, 252, 431–434. [Google Scholar] [CrossRef] [PubMed]
  164. Harvey, B.; Hackett, N.R.; El-sawy, T.; Rosengart, T.K.; Hirschowitz, E.A.; Lieberman, M.D.; Lesser, M.L.; Crystal, R.G.; Al, H.E.T. Variability of Human Systemic Humoral Immune Responses to Adenovirus Gene Transfer Vectors Administered to Different Organs. J. Virol. 1999, 73, 6729–6742. [Google Scholar] [CrossRef] [Green Version]
  165. Crystal, R.G. Adenovirus: The First Effective in Vivo Gene Delivery Vector. Hum. Gene Ther. 2014, 25, 3–11. [Google Scholar] [CrossRef] [Green Version]
  166. Raper, S.E.; Chirmule, N.; Lee, F.S.; Wivel, N.A.; Bagg, A.; Gao, G.P.; Wilson, J.M.; Batshaw, M.L. Fatal Systemic Inflammatory Response Syndrome in a Ornithine Transcarbamylase Deficient Patient Following Adenoviral Gene Transfer. Mol. Genet. Metab. 2003, 80, 148–158. [Google Scholar] [CrossRef]
  167. Fang, B.; Eisensmith, R.C.; Wang, H.; Kay, M.A.; Cross, R.E.; Landen, C.N.; Gordon, G.; Bellinger, D.A.; Read, M.S.; Hu, P.C.; et al. Gene Therpy for Hemophilia B: Host Prolongs the Therapeutic Effect o f Factor I X Immunosuppression. Hum. Gene Ther. 1995, 1044, 1039–1044. [Google Scholar] [CrossRef]
  168. A Smith, T.; White, B.D.; Gardner, J.M.; Kaleko, M.; McClelland, A. Transient Immunosuppression Permits Successful Repetitive Intravenous Administration of an Adenovirus Vector. Gene Ther. 1996, 3, 496–502. [Google Scholar] [PubMed]
  169. Goulet, M.A.R.L.Y.N.E.; Gravel, C.; Roy, R.; Tremblay, J.P. Immunosuppression to Control the Immune Reactions Triggered by First-Generation Gene Transfer. Hum. Gene Ther. 1995, 1401, 1391–1401. [Google Scholar]
  170. Poller, W.; Schneider-Rasp, S.; Liebert, U.; Merklein, F.; Thalheimer, P.; Haack, A.; Schwaab, R.; Schmitt, C.; Brackmann, H.H. Brackmann’ Stabilization of Transgene Expression by Incorporation of E3 Region Genes into an Adenoviral Factor IX Vector and by Transient Anti-CD4 Treatment of the Host. Gene Ther. 1996, 3, 521–530. [Google Scholar] [PubMed]
  171. Sawchuk, S.J.; Boivin, G.P.; Duwel, L.E.; Ball, W.; Bove, K.; Trapnell, B.; Hirsch, R.; Al, S.E.T. Anti-T Cell Receptor Monoclonal Antibody Transgene Expression Following Gene Prolongs Vivo Adenovirus-Mediated In Synovium Transfer to Mouse Synovium. Hum Gene Ther. 1996, 7, 499–506. [Google Scholar] [CrossRef] [PubMed]
  172. Cassivi, S.D.; Liu, M.; Boehler, A.; Tanswell, A.K.; Slutsky, A.S.; Keshavjee, S.; Wechsler, A.S.; Rosengart, T.; Carpentier, A.F.; Robbins, R.C. Transgene Expression after Adenovirus-Mediated Retransfection of Rat Lungs Is Increased and Prolonged by Transplant Immunosuppression. J. Thorac. Cardiovasc. Surg. 1999, 117, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Wang, H.; Li, Z.; Yumul, R.; Lara, S.; Hemminki, A.; Fender, P.; Lieber, A. Multimerization of Adenovirus Serotype 3 Fiber Knob Domains Is Required for Efficient Binding of Virus to Desmoglein 2 and Subsequent Opening of Epithelial Junctions. J. Virol. 2011, 85, 6390–6402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Gaggar, A.; Shayakhmetov, D.M.; Lieber, A. CD46 Is a Cellular Receptor for Group B Adenoviruses. Nat. Med. 2003, 9, 1408–1412. [Google Scholar] [CrossRef]
  175. Nilsson, M.; Karlsson, S.; Fan, X. Functionally Distinct Subpopulations of Cord Blood CD34+ Cells Are Transduced by Adenoviral Vectors with Serotype 5 or 35 Tropism. Mol. Ther. 2004, 9, 377–388. [Google Scholar] [CrossRef]
  176. Shayakhmetov, D.M.; Papayannopoulou, T.; Stamatoyannopoulos, G.; Lieber, A. Efficient Gene Transfer into Human CD34+ Cells by a Retargeted Adenovirus Vector. J. Virol. 2000, 74, 2567–2583. [Google Scholar] [CrossRef] [Green Version]
  177. Ni, S.; Bernt, K.; Gaggar, A.; Li, Z.Y.; Kiem, H.P.; Lieber, A. Evaluation of Biodistribution and Safety of Adenovirus Vectors Containing Group B Fibers after Intravenous Injection into Baboons. Hum. Gene Ther. 2005, 16, 664–677. [Google Scholar] [CrossRef] [Green Version]
  178. Alba, R.; Bosch, A.; Chillon, M. Gutless Adenovirus: Last-Generation Adenovirus for Gene Therapy. Gene Ther. 2005, 12, S18–S27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Ricobaraza, A.; Gonzalez-Aparicio, M.; Mora-Jimenez, L.; Lumbreras, S.; Hernandez-Alcoceba, R. High-Capacity Adenoviral Vectors: Expanding the Scope of Gene Therapy. Int. J. Mol. Sci. 2020, 21, 3643. [Google Scholar] [CrossRef] [PubMed]
  180. Sandoval-Villegas, N.; Nurieva, W.; Amberger, M.; Ivics, Z. Contemporary Transposon Tools: A Review and Guide through Mechanisms and Applications of Sleeping Beauty, Piggybac and Tol2 for Genome Engineering. Int. J. Mol. Sci. 2021, 22, 5084. [Google Scholar] [CrossRef] [PubMed]
  181. Smith, R.P.; Riordan, J.D.; Feddersen, C.R.; Dupuy, A.J. A Hybrid Adenoviral Vector System Achieves Efficient Long-Term Gene Expression in the Liver via PiggyBac Transposition. Hum. Gene Ther. 2015, 26, 377–385. [Google Scholar] [CrossRef] [Green Version]
  182. Richter, M.; Saydaminova, K.; Yumul, R.; Krishnan, R.; Liu, J.; Nagy, E.E.; Singh, M.; Izsvák, Z.; Cattaneo, R.; Uckert, W.; et al. In Vivo Transduction of Primitive Mobilized Hematopoietic Stem Cells after Intravenous Injection of Integrating Adenovirus Vectors. Blood 2016, 128, 2206–2217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Wang, H.; Richter, M.; Psatha, N.; Li, C.; Kim, J.; Liu, J.; Ehrhardt, A.; Nilsson, S.K.; Cao, B.; Palmer, D.; et al. A Combined in Vivo HSC Transduction/Selection Approach Results in Efficient and Stable Gene Expression in Peripheral Blood Cells in Mice. Mol. Ther.-Methods Clin. Dev. 2018, 8, 52–64. [Google Scholar] [CrossRef] [Green Version]
  184. Psatha, N.; Sgouramali, E.; Gkountis, A.; Siametis, A.; Baliakas, P.; Constantinou, V.; Athanasiou, E.; Arsenakis, M.; Anagnostopoulos, A.; Papayannopoulou, T.; et al. Superior Long-Term Repopulating Capacity of G-CSF + Plerixafor-Mobilized Blood: Implications for Stem Cell Gene Therapy by Studies in the Hbbth-3 Mouse Model. Hum. Gene Ther. Methods 2014, 25, 317–327. [Google Scholar] [CrossRef] [Green Version]
  185. Yannaki, E.; Karponi, G.; Zervou, F.; Constantinou, V.; Bouinta, A.; Tachynopoulou, V.; Kotta, K.; Jonlin, E.; Papayannopoulou, T.; Anagnostopoulos, A.; et al. Hematopoietic Stem Cell Mobilization for Gene Therapy: Superior Mobilization by the Combination of Granulocyte-Colony Stimulating Factor plus Plerixafor in Patients with β-Thalassemia Major. Hum. Gene Ther. 2013, 24, 852–860. [Google Scholar] [CrossRef] [Green Version]
  186. Psatha, N.; Yannaki, E.; Athanasiou, E.; Sgouramalli, E.; Mantenoudi, O.; Arsenakis, M.; Anagnostopoulos, A.; Fassas, A. The Combination of AMD3100+G-CSF Successfully Mobilizes HSCs into the Peripheral Blood Compared to G-CSF Alone, in a Thalassemic Mouse Model. Haematol. Hematol. J. 2010, 95, 447. [Google Scholar]
  187. Karponi, G.; Psatha, N.; Lederer, C.W.; Adair, J.E.; Zervou, F.; Zogas, N.; Kleanthous, M.; Tsatalas, C.; Anagnostopoulos, A.; Sadelain, M.; et al. Plerixafor+G-CSF-Mobilized CD34+ Cells Represent an Optimal Graft Source for Thalassemia Gene Therapy. Blood 2015, 126, 616–619. [Google Scholar] [CrossRef] [Green Version]
  188. Diana, J.; Manceau, S.; Leblanc, T.; Magnani, A.; Magrin, E.; Bendavid, M.; Couzin, C.; Joseph, L.; Soulier, J.; Cavazzana, M.; et al. A New Step in Understanding Stem Cell Mobilization in Patients with Fanconi Anemia: A Bridge to Gene Therapy. Transfussion 2021, 62, 165–172. [Google Scholar] [CrossRef] [PubMed]
  189. Hsieh, M.M.; Tisdale, J.F. Hematopoietic Stem Cell Mobilization with Plerixafor in Sickle Cell Disease. Haematologica 2018, 103, 749–750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Uchida, N.; Leonard, A.; Stroncek, D.; Panch, S.R.; West, K.; Molloy, E.; Hughes, T.E.; Hauffe, S.; Taylor, T.; Fitzhugh, C.; et al. Safe and Efficient Peripheral Blood Stem Cell Collection in Patients with Sickle Cell Disease Using Plerixafor. Hematologica 2020, 105, e497. [Google Scholar] [CrossRef]
  191. Wang, H.; Liu, Z.; Li, C.; Gil, S.; Papayannopoulou, T.; Doering, C.B.; Lieber, A. High-Level Protein Production in Erythroid Cells Derived from in Vivo Transduced Hematopoietic Stem Cells. Blood Adv. 2019, 3, 2883–2894. [Google Scholar] [CrossRef] [PubMed]
  192. Humbert, O.; Chan, F.; Rajawat, Y.S.; Torgerson, T.R.; Burtner, C.R.; Hubbard, N.W.; Humphrys, D.; Norgaard, Z.K.; O’Donnell, P.; Adair, J.E.; et al. Rapid Immune Reconstitution of SCID-X1 Canines after G-CSF/AMD3100 Mobilization and in Vivo Gene Therapy. Blood Adv. 2018, 2, 987–999. [Google Scholar] [CrossRef] [Green Version]
  193. Li, C.; Psatha, N.; Wang, H.; Singh, M.; Samal, H.B.; Zhang, W.; Ehrhardt, A.; Izsvák, Z.; Papayannopoulou, T.; Lieber, A. Integrating HDAd5/35++ Vectors as a New Platform for HSC Gene Therapy of Hemoglobinopathies. Mol. Ther.-Methods Clin. Dev. 2018, 9, 142–152. [Google Scholar] [CrossRef] [Green Version]
  194. Wang, H.; Georgakopoulou, A.; Psatha, N.; Li, C.; Capsali, C.; Samal, H.B.; Anagnostopoulos, A.; Ehrhardt, A.; Izsvák, Z.; Papayannopoulou, T.; et al. In Vivo Hematopoietic Stem Cell Gene Therapy Ameliorates Murine Thalassemia Intermedia. J. Clin. Invest. 2019, 129, 598–615. [Google Scholar] [CrossRef]
  195. Wang, H.; Georgakopoulou, A.; Li, C.; Liu, Z.; Gil, S.; Bashyam, A.; Yannaki, E.; Anagnostopoulos, A.; Pande, A.; Izsvák, Z.; et al. Curative in Vivo Hematopoietic Stem Cell Gene Therapy of Murine Thalassemia Using Large Regulatory Elements. JCI Insight 2020, 5, e139538. [Google Scholar] [CrossRef]
  196. Wang, M.Y.; Zhao, R.; Gao, L.J.; Gao, X.F.; Wang, D.P.; Cao, J.M. SARS-CoV-2: Structure, Biology, and Structure-Based Therapeutics Development. Front. Cell. Infect. Microbiol. 2020, 10, 587269. [Google Scholar] [CrossRef]
  197. Wang, H.; Li, C.; Obadan, A.O.; Frizzell, H.; Hsiang, T.-Y.; Gil, S.; Germond, A.; Fountain, C.; Baldessari, A.; Roffler, S.; et al. In Vivo Hematopoietic Stem Cell Gene Therapy for SARS-CoV2 Infection Using a Decoy Receptor. Hum. Gene Ther. 2022, 33, 389–403. [Google Scholar] [CrossRef]
  198. Saydaminova, K.; Ye, X.; Wang, H.; Richter, M.; Ho, M.; Chen, H.Z.; Xu, N.; Kim, J.S.; Papapetrou, E.; Holmes, M.C.; et al. Efficient Genome Editing in Hematopoietic Stem Cells with Helper-Dependent Ad5/35 Vectors Expressing Site-Specific Endonucleases under MicroRNA Regulation. Mol. Ther.-Methods Clin. Dev. 2015, 2, 14057. [Google Scholar] [CrossRef] [PubMed]
  199. Li, C.; Psatha, N.; Sova, P.; Gil, S.; Wang, H.; Kim, J.; Kulkarni, C.; Valensisi, C.; David Hawkins, R.; Stamatoyannopoulos, G.; et al. Reactivation of G-Globin in Adult b-YAC Mice after Ex Vivo and in Vivo Hematopoietic Stem Cell Genome Editing. Blood 2018, 131, 2915–2928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Li, C.; Mishra, A.S.; Gil, S.; Wang, M.; Georgakopoulou, A.; Papayannopoulou, T.; Hawkins, R.D.; Lieber, A. Targeted Integration and High-Level Transgene Expression in AAVS1 Transgenic Mice after In Vivo HSC Transduction with HDAd5/35++ Vectors. Mol. Ther. 2019, 27, 2195–2212. [Google Scholar] [CrossRef] [PubMed]
  201. Li, C.; Wang, H.; Gil, S.; Germond, A.; Fountain, C.; Baldessari, A.; Kim, J.; Liu, Z.; Georgakopoulou, A.; Radtke, S.; et al. Safe and Efficient in Vivo Hematopoietic Stem Cell Transduction in Nonhuman Primates Using HDAd5/35++ Vectors. Mol. Ther.-Methods Clin. Dev. 2022, 24, 127–141. [Google Scholar] [CrossRef]
  202. Chew, W.L.; Tabebordbar, M.; Cheng, J.K.; Mali, P.; Wu, E.Y.; Ng, A.H.; Zhu, K.; Wagers, A.J.; Church, G.M. A Multi-Functional AAV-CRISPR-Cas9 and Its Host Response. Nat. Methods 2016, 13, 868–874. [Google Scholar] [CrossRef] [Green Version]
  203. Vandamme, C.; Xicluna, R.; Hesnard, L.; Devaux, M.; Jaulin, N.; Guilbaud, M.; Le Duff, J.; Couzinié, C.; Moullier, P.; Saulquin, X.; et al. Tetramer-Based Enrichment of Preexisting Anti-AAV8 CD8+ T Cells in Human Donors Allows the Detection of a TEMRA Subpopulation. Front. Immunol. 2020, 10, 3110. [Google Scholar] [CrossRef]
  204. Ferla, R.; Alliegro, M.; Dell’Anno, M.; Nusco, E.; Cullen, J.M.; Smith, S.N.; Wolfsberg, T.G.; O’Donnell, P.; Wang, P.; Nguyen, A.D.; et al. Low Incidence of Hepatocellular Carcinoma in Mice and Cats Treated with Systemic Adeno-Associated Viral Vectors. Mol. Ther.-Methods Clin. Dev. 2021, 20, 247–257. [Google Scholar] [CrossRef]
  205. Colella, P.; Ronzitti, G.; Mingozzi, F. Emerging Issues in AAV-Mediated In Vivo Gene Therapy. Mol. Ther.-Methods Clin. Dev. 2018, 8, 87–104. [Google Scholar] [CrossRef] [Green Version]
  206. Wu, P.; Chen, H.; Jin, R.; Weng, T.; Ho, J.K.; You, C.; Zhang, L.; Wang, X.; Han, C. Non-Viral Gene Delivery Systems for Tissue Repair and Regeneration. J. Transl. Med. 2018, 16, 29. [Google Scholar] [CrossRef] [Green Version]
  207. Cordeiro, R.A.; Santo, D.; Farinha, D.; Serra, A.; Faneca, H.; Coelho, J.F.J. High Transfection Efficiency Promoted by Tailor-Made Cationic Tri-Block Copolymer-Based Nanoparticles. Acta Biomater. 2017, 47, 113–123. [Google Scholar] [CrossRef]
  208. Jung, M.R.; Shim, I.K.; Kim, E.S.; Park, Y.J.; Yang, Y.I.; Lee, S.K.; Lee, S.J. Controlled Release of Cell-Permeable Gene Complex from Poly(L-Lactide) Scaffold for Enhanced Stem Cell Tissue Engineering. J. Control. Release 2011, 152, 294–302. [Google Scholar] [CrossRef] [PubMed]
  209. Islam, M.A.; Park, T.E.; Singh, B.; Maharjan, S.; Firdous, J.; Cho, M.H.; Kang, S.K.; Yun, C.H.; Choi, Y.J.; Cho, C.S. Major Degradable Polycations as Carriers for DNA and SiRNA. J. Control. Release 2014, 193, 74–89. [Google Scholar] [CrossRef] [PubMed]
  210. Majidi, A.; Nikkhah, M.; Sadeghian, F.; Hosseinkhani, S. Development of Novel Recombinant Biomimetic Chimeric MPG-Based Peptide as Nanocarriers for Gene Delivery: Imitation of a Real Cargo. Eur. J. Pharm. Biopharm. 2016, 107, 191–204. [Google Scholar] [CrossRef]
  211. Fernández, E.F.; Santos-Carballal, B.; de Santi, C.; Ramsey, J.M.; MacLoughlin, R.; Cryan, S.A.; Greene, C.M. Biopolymer-Based Nanoparticles for Cystic Fibrosis Lung Gene Therapy Studies. Materials 2018, 11, 122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Hyde, S.C.; Southern, K.W.; Gileadi, U.; Fitzjohn, E.M.; Mofford, K.A.; Waddell, B.E.; Gooi, H.C.; Goddard, C.A.; Hannavy, K.; Smyth, S.E.; et al. Repeat Administration of DNA/Liposomes to the Nasal Epithelium of Patients with Cystic Fibrosis. Gene Ther. 2000, 7, 1156–1165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Alton, E.W.F.W.; Armstrong, D.K.; Ashby, D.; Bayfield, K.J.; Bilton, D.; Bloomfield, E.V.; Boyd, A.C.; Brand, J.; Buchan, R.; Calcedo, R.; et al. Repeated Nebulisation of Non-Viral CFTR Gene Therapy in Patients with Cystic Fibrosis: A Randomised, Double-Blind, Placebo-Controlled, Phase 2b Trial. Lancet Respir. Med. 2015, 3, 684–691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Sato, Y.; Matsui, H.; Yamamoto, N.; Sato, R.; Munakata, T.; Kohara, M.; Harashima, H. Highly Specific Delivery of SiRNA to Hepatocytes Circumvents Endothelial Cell-Mediated Lipid Nanoparticle-Associated Toxicity Leading to the Safe and Efficacious Decrease in the Hepatitis B Virus. J. Control. Release 2017, 266, 216–225. [Google Scholar] [CrossRef] [PubMed]
  215. Böttger, R.; Pauli, G.; Chao, P.H.; AL Fayez, N.; Hohenwarter, L.; Li, S.D. Lipid-Based Nanoparticle Technologies for Liver Targeting. Adv. Drug Deliv. Rev. 2020, 154–155, 79–101. [Google Scholar] [CrossRef]
  216. 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]
  217. McNeer, N.A.; Schleifman, E.B.; Cuthbert, A.; Brehm, M.; Jackson, A.; Cheng, C.; Anandalingam, K.; Kumar, P.; Shultz, L.D.; Greiner, D.L.; et al. Systemic Delivery of Triplex-Forming PNA and Donor DNA by Nanoparticles Mediates Site-Specific Genome Editing of Human Hematopoietic Cells in Vivo. Gene Ther. 2013, 20, 658–669. [Google Scholar] [CrossRef] [Green Version]
  218. Bahal, R.; Ali McNeer, N.; Quijano, E.; Liu, Y.; Sulkowski, P.; Turchick, A.; Lu, Y.C.; Bhunia, D.C.; Manna, A.; Greiner, D.L.; et al. In Vivo Correction of Anaemia in β-Thalassemic Mice by Γ3PNA-Mediated Gene Editing with Nanoparticle Delivery. Nat. Commun. 2016, 7, 13304. [Google Scholar] [CrossRef] [PubMed]
  219. Cruz, L.J.; van Dijk, T.; Vepris, O.; Li, T.M.W.Y.; Schomann, T.; Baldazzi, F.; Kurita, R.; Nakamura, Y.; Grosveld, F.; Philipsen, S.; et al. PLGA-Nanoparticles for Intracellular Delivery of the CRISPR-Complex to Elevate Fetal Globin Expression in Erythroid Cells. Biomaterials 2021, 268, 120580. [Google Scholar] [CrossRef] [PubMed]
  220. Cannon, P.; Asokan, A.; Czechowicz, A.; Hammond, P.; Kohn, D.B.; Lieber, A.; Malik, P.; Marks, P.; Porteus, M.; Verhoeyen, E.; et al. Safe and Effective in Vivo Targeting and Gene Editing in Hematopoietic Stem Cells: Strategies for Accelerating Development. Hum. Gene Ther. 2021, 32, 31–42. [Google Scholar] [CrossRef] [PubMed]
  221. Charlesworth, C.T.; Deshpande, P.S.; Dever, D.P.; Camarena, J.; Lemgart, V.T.; Cromer, M.K.; Vakulskas, C.A.; Collingwood, M.A.; Zhang, L.; Bode, N.M.; et al. Identification of Preexisting Adaptive Immunity to Cas9 Proteins in Humans. Nat. Med. 2019, 25, 249–254. [Google Scholar] [CrossRef]
  222. Simhadri, V.L.; McGill, J.; McMahon, S.; Wang, J.; Jiang, H.; Sauna, Z.E. Prevalence of Pre-Existing Antibodies to CRISPR-Associated Nuclease Cas9 in the USA Population. Mol. Ther.-Methods Clin. Dev. 2018, 10, 105–112. [Google Scholar] [CrossRef]
  223. Wagner, D.L.; Amini, L.; Wendering, D.J.; Burkhardt, L.M.; Akyüz, L.; Reinke, P.; Volk, H.D.; Schmueck-Henneresse, M. High Prevalence of Streptococcus Pyogenes Cas9-Reactive T Cells within the Adult Human Population. Nat. Med. 2019, 25, 242–248. [Google Scholar] [CrossRef]
  224. Ferdosi, S.R.; Ewaisha, R.; Moghadam, F.; Krishna, S.; Park, J.G.; Ebrahimkhani, M.R.; Kiani, S.; Anderson, K.S. Multifunctional CRISPR-Cas9 with Engineered Immunosilenced Human T Cell Epitopes. Nat. Commun. 2019, 10, 1842. [Google Scholar] [CrossRef] [Green Version]
  225. Li, A.; Tanner, M.R.; Lee, C.M.; Hurley, A.E.; De Giorgi, M.; Jarrett, K.E.; Davis, T.H.; Doerfler, A.M.; Bao, G.; Beeton, C.; et al. AAV-CRISPR Gene Editing Is Negated by Pre-Existing Immunity to Cas9. Mol. Ther. 2020, 28, 1432–1441. [Google Scholar] [CrossRef]
  226. Monteilhet, V.; Veron, P.; Leborgne, C.; Benveniste, O. Prevalence of Serum IgG and Neutralizing Factors Against Adeno-Associated Virus (AAV) Types 1, 2, 5, 6, 8, and 9 in the Healthy Population: Implications for Gene Therapy Using AAV Vectors. Hum. Gene Ther. 2010, 712, 704–712. [Google Scholar]
  227. Reddy, P.S.; Ganesh, S.; Limbach, M.P.; Brann, T.; Pinkstaff, A.; Kaloss, M.; Kaleko, M.; Connelly, S. Development of Adenovirus Serotype 35 as a Gene Transfer Vector. Virology 2003, 311, 384–393. [Google Scholar] [CrossRef] [Green Version]
  228. Nathwani, A.C.; Reiss, U.M.; Tuddenham, E.G.D.; Rosales, C.; Chowdary, P.; McIntosh, J.; Della Peruta, M.; Lheriteau, E.; Patel, N.; Raj, D.; et al. Long-Term Safety and Efficacy of Factor IX Gene Therapy in Hemophilia B. N. Engl. J. Med. 2014, 371, 1994–2004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Chowdary, P.; Shapiro, S.; Makris, M.; Evans, G.; Boyce, S.; Talks, K.; Dolan, G.; Reiss, U.; Phillips, M.; Riddell, A.; et al. Phase 1-2 Trial of AAVS3 Gene Therapy in Patients with Hemophilia B. N. Engl. J. Med. 2022, 387, 237–247. [Google Scholar] [CrossRef] [PubMed]
  230. Koo, T.; Lee, J.; Kim, J.S. Measuring and Reducing Off-Target Activities of Programmable Nucleases Including CRISPR-Cas9. Mol. Cells 2015, 38, 475–481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Shengdar, Q.; Tsai, J.K.J. Defining and Improving the Genome-Wide Specificities of CRISPR-Cas9 Nucleases. Nat. Rev. Genet. 2016, 17, 300–312. [Google Scholar] [CrossRef]
  232. Akcakaya, P.; Bobbin, M.L.; Guo, J.A.; Malagon-lopez, J.; Clement, K.; Garcia, S.P.; Fellows, M.D.; Porritt, M.J.; Firth, M.A.; Carreras, A.; et al. In Vivo CRISPR Editing with No Detectable Genome-Wide Offtarget Mutations. Nature 2018, 561, 416–419. [Google Scholar] [CrossRef]
  233. Atkins, A.; Chung, C.-H.; Allen, A.G.; Dampier, W.; Gurrola, T.E.; Sariyer, I.K.; Nonnemacher, M.R.; Wigdahl, B. Off-Target Analysis in Gene Editing and Applications for Clinical Translation of CRISPR/Cas9 in HIV-1 Therapy. Front. Genome Ed. 2021, 3, 673022. [Google Scholar] [CrossRef]
  234. Kosicki, M.; Tomberg, K.; Bradley, A. Repair of Double-Strand Breaks Induced by CRISPR–Cas9 Leads to Large Deletions and Complex Rearrangements. Nat. Biotechnol. 2018, 36, 765–771. [Google Scholar] [CrossRef]
  235. Nahmad, A.D.; Reuveni, E.; Goldschmidt, E.; Tenne, T.; Liberman, M.; Horovitz-Fried, M.; Khosravi, R.; Kobo, H.; Reinstein, E.; Madi, A.; et al. Frequent Aneuploidy in Primary Human T Cells after CRISPR-Cas9 Cleavage. Nat. Biotechnol. 2022. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Psatha, N.; Paschoudi, K.; Papadopoulou, A.; Yannaki, E. In Vivo Hematopoietic Stem Cell Genome Editing: Perspectives and Limitations. Genes 2022, 13, 2222. https://doi.org/10.3390/genes13122222

AMA Style

Psatha N, Paschoudi K, Papadopoulou A, Yannaki E. In Vivo Hematopoietic Stem Cell Genome Editing: Perspectives and Limitations. Genes. 2022; 13(12):2222. https://doi.org/10.3390/genes13122222

Chicago/Turabian Style

Psatha, Nikoletta, Kiriaki Paschoudi, Anastasia Papadopoulou, and Evangelia Yannaki. 2022. "In Vivo Hematopoietic Stem Cell Genome Editing: Perspectives and Limitations" Genes 13, no. 12: 2222. https://doi.org/10.3390/genes13122222

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