Next Article in Journal
PBAF Subunit Pbrm1 Selectively Influences the Transition from Progenitors to Pre-Myelinating Cells during Oligodendrocyte Development
Previous Article in Journal
Generation of Red Blood Cells from Human Pluripotent Stem Cells—An Update
Previous Article in Special Issue
CRISPR-Cas System: The Current and Emerging Translational Landscape
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genome Editing for Cystic Fibrosis

Department of Microbiology, Immunology and Parasitology, Louisiana State University Health Sciences Center, CSRB 607, 533 Bolivar Street, New Orleans, LA 70112, USA
Cells 2023, 12(12), 1555; https://doi.org/10.3390/cells12121555
Submission received: 1 April 2023 / Revised: 6 May 2023 / Accepted: 1 June 2023 / Published: 6 June 2023
(This article belongs to the Special Issue CRISPR-Based Genome Editing in Translational Research)

Abstract

:
Cystic fibrosis (CF) is a monogenic recessive genetic disorder caused by mutations in the CF Transmembrane-conductance Regulator gene (CFTR). Remarkable progress in basic research has led to the discovery of highly effective CFTR modulators. Now ~90% of CF patients are treatable. However, these modulator therapies are not curative and do not cover the full spectrum of CFTR mutations. Thus, there is a continued need to develop a complete and durable therapy that can treat all CF patients once and for all. As CF is a genetic disease, the ultimate therapy would be in-situ repair of the genetic lesions in the genome. Within the past few years, new technologies, such as CRISPR/Cas gene editing, have emerged as an appealing platform to revise the genome, ushering in a new era of genetic therapy. This review provided an update on this rapidly evolving field and the status of adapting the technology for CF therapy.

1. Introduction

CRISPR—clustered regularly interspaced short palindromic repeats—were first discovered in the 3′-end flanking region of the iap gene in E. coli in 1987 [1] and later found in many other bacteria and even archaea [2,3,4]. These repeats and their associated proteins constitute an adaptive immune system to protect host microbes from invasion by foreign genetic elements, such as viruses, through targeted cleavage of the nucleic acids with RNA-guided endonuclease [5,6,7,8]. Among many types of CRISPR-associated proteins (Cas), Streptococcus pyogenes Cas9 (SpCas9) is the most studied [9,10]. Jennifer Doudna, Emmanuelle Charpentier, and their teams first engineered the CRISPR/Cas9 system for programmable gene editing in 2012 [11]. Then in the following year, Feng Zhang, George Church, and their colleagues published the first eukaryotic genome engineering with the system [12,13], which opened the floodgates for genome engineering in living cells of diverse species. This breakthrough has led to revolutionary advances in medicine through the direct correction of disease-associated genes in the genome for therapy. To date, a decade after the invention of programmable gene editing, there have been 76 clinical trials registered with the US Food and Drug Administration (ClinicalTrials.gov) to test CRISPR/Cas technology in different diseases or conditions. Excitingly, the first such trial for sickle cell disease and β-thalassemia achieved a sustained cure for these two deadly transfusion-dependent blood diseases [14], greatly inspiring basic researchers and clinical practitioners to pursue similar treatments for other genetic diseases, including cystic fibrosis (CF).

2. CF Gene Mutations and CF Clinical Diseases

2.1. CF Genomic Mutations

CF is one of the most common life-threatening genetic disorders, affecting all races and ethnic groups, but the Caucasian population has the highest rate of incidence [15]. There are currently over 100,000 CF patients worldwide [16]. CF genetic lesions have been mapped to human chromosome 7 (7q31.2) in 1989 by Francis Collins and Lap-chee Tsui’s groups [17,18]. The normal function of the CF locus is to direct the synthesis of CF transmembrane-conductance regulator (CFTR) protein, a cAMP-activated anion channel [19]. Since the discovery of CFTR [18,20], more than 2000 mutations have been identified, which cause clinical illnesses with varying severity [21,22]. Based on CFTR protein production and channel activity, CF mutations can be categorized into five classes [23,24]. Class 1 contains nonsense mutations, e.g., G542X, giving rise to premature termination of CFTR protein. Class 2, represented by F508del, produces a full-length, but misfolded CFTR protein that degrades in the endoplasmic reticulum. Class 3, such as G551D, results in a full-length CFTR with a null channel function. Class 4 leads to a mature, correctly targeted CFTR with partial channel function. Class 5 produces insufficient CFTR protein. The F508del mutation in one or two alleles occupies ~85.5% of all CF individuals, and the other CFTR mutations take up the remaining ~14.5% [25].

2.2. CF Clinical Diseases

Clinically, CF presents as a progressive, chronic, and debilitating illness [26] that affects multiple organ systems, including the lung, sinuses, gastrointestinal tract, liver, pancreas, and vas deferens. Among all these afflicted organ systems, CF lung disease is predominant in the adult patient population, accounting for over 90% of CF morbidity and mortality. CFTR transports chloride and bicarbonate anions across the apical membrane of airway epithelia and regulates sodium absorption and airway surface liquid hydration [27]. CFTR dysfunction impairs anion secretion, increases sodium absorption, causes airway surface liquid dehydration, and disables mucociliary clearance [28]. It is widely believed that these conditions induce chronic bacterial infection, persistent neutrophilic inflammation, and small airway mucopurulent obstruction, which ultimately cause bronchiectasis and pulmonary failure.
The pancreas is one of the earliest- and most commonly-affected organs in people with CF. The term “cystic fibrosis” was named after the fibrocystic lesions observed in the pancreas from pediatric autopsy cases [29]. Ductal obstruction leads to pancreatic protein buildup behind the obstructive sites, which induces inflammation, fibrosis, fatty replacement, and eventually pancreatic destruction. Approximately 85% of CF patients become pancreatic insufficient, requiring lifelong pancreatic enzyme replacement [30]. CF-related diabetes is an increasingly recognized complication of CF that occurs in 40–50% of adults with CF [31].
CF intestinal disease begins in the uterus and is often the first sign of CF encountered clinically. Meconium ileus (MI) presents in up to 20% of neonates with CF [32]. The inspissated meconium obstructs the small intestine at the terminal ileum. If left untreated, the prognosis is poor. In infancy and childhood, CF patients demonstrate nutrient malabsorption, steatorrhea, and inadequate weight and body length, for which pancreatic insufficiency is largely responsible [33]. Pancreatic enzyme replacement therapy proves beneficial in overcoming some of these problems. However, distal intestinal obstruction syndrome (DIOS) and constipation continue to be a problem throughout the patients’ lives [34]. Adult CF patients demonstrate small intestine bacterial overgrowth (SIBO) [35] and large intestine dysbiosis [36,37]. CF intestinal disease is also marked by a swollen intestinal wall, intestinal stricture, excessive neutrophil infiltration, and fibrosing colonopathy [38,39,40,41,42,43,44].
CF liver disease (CFLD) affects ~30% of CF patients. Clinical manifestations of CFLD are elevation of serum liver enzymes, hepatic steatosis, focal biliary cirrhosis, multilobular biliary cirrhosis, neonatal cholestasis, cholelithiasis, cholecystitis, and micro-gallbladder [45]. CFTR chloride channel is expressed to the apical membrane of cholangiocytes lining the biliary ducts [46], essential to maintain pH regulation and biliary HCO3 secretion. A proper alkaline balance is believed to protect cholangiocytes against hydrophobic bile acids [47].
CF damages the reproductive system, with a majority of adult male patients showing congenital bilateral absence of vas deferens [48]. Females with CF are found less fertile than normal healthy women, and some show congenital absence of the uterus and vagina [49].
The CFTR channel is also expressed in many non-epithelial cells, including endothelial cells [50], skeletal muscle [51], smooth muscle [52,53], cardiac muscle [54], red blood cells [55], platelets [56], neurons [57] and even reproductive sperm cells [58], suggesting a role of CFTR in the functions of these cells. More importantly, immune cells, such as neutrophils and monocytes/macrophages, have CFTR expression [59,60,61]. As infection and inflammation are the two major pathologies in CF, it was predicted that the host immune system is impaired. Indeed, CFTR is a major chloride channel that transports chloride to phagosomes to produce hypochlorous acid, a potent antimicrobial oxidant [61,62,63,64,65,66]. CFTR loss of function in neutrophils impairs the host’s defense against selective pathogens and compromises the host’s ability to resolve inflammation [59,67,68,69,70]. Other abnormalities are also reported in CF neutrophils, including suboptimal activation [71], cleavage of CXCR1 [72], hyper-sensitivity to LPS stimulation [73], deviant production of reactive oxygen species [74], genome-wide gene expression perturbation [75], alteration in inflammatory signaling [76], hyper-production of IL-8 [77,78], delayed apoptosis [79], abnormal extracellular trap formation [80], hyper-oxidation of glutathione [81], and lately abnormal granule release [82]. Moreover, CF monocytes/macrophages exhibit deficiencies in their host defense and other functions, including hyper-sensitivity to challenge [83,84,85], impaired capacity of killing internalized bacteria [60,86,87], and reduced scavenger ability [88]. In addition to the infection and inflammation complications, CF patients often suffer arthropathies [89,90]. CF-related arthropathy has a rate of occurrence ranging from 2% to 8.5% [91] with distinct symptoms, such as recurrent episodes of joint pain, swelling, tenderness, and limitation of movement. One or more joints may be affected, and there may also be fever and skin manifestations [90]. Moreover, 2–7% of CF patients suffer hypertrophic pulmonary osteoarthropathy with long-bone periostitis, and finger and toe clubbing with symmetric arthralgia [92,93]. These inflammation-related clinical manifestations further demonstrate the CF immune defect.

3. General Considerations in Gene Editing Design for CF

First, CF is a systemic disease that affects many epithelial and non-epithelial organs. What is the best route to deliver gene editing? From a therapeutic perspective, the ideal way is through systemic delivery that can reach all tissues with one application. Second, CF has more than 2000 distinct mutations. How can gene editing achieve pan-mutation correction with one formula? One strategy is to perform targeted insertion of a full-length CFTR cDNA into the CFTR locus to replace the endogenous CFTR. This one-size-fits-all approach can target almost all CF mutations. Third, terminally differentiated epithelial cells have a certain life span. How can durable gene editing be achieved? Stem and progenitor cells would be the preferential targets for correction. Fourth, CF clinical diseases develop early in life. What is the best age to perform gene editing for therapy? It is apparently important to target CF before damage to organ structures and functions become irreversible. Thus, gene editing components and formulation must be applicable to young patients. Fifth, safe genome manipulation is paramount. Efforts to minimize off-target effects and to ensure long-term safety should guide any gene editing design. Attention should be specially paid to the systemic delivery of gene editing to young patients, as germline modification may occur.

4. Gene Editing Tools for Selection

4.1. Cas Nuclease Editors

There are 2 classes of CRISPR/Cas systems, which can be subdivided into 6 types and at least 29 subtypes according to their structure and function of Cas protein [94]. Class I contains Types I, III and IV, while Class 2 contains Types II, V and VI. Each member in Class I has multiple effector complexes with several Cas proteins, whereas each Class II member has a single, large, multidomain Cas protein. Due to the relatively simple ribonucleoprotein complexes involved, Class II members have been extensively studied and are the most widely used tool for genome engineering.
The prototype CRISPR/Cas9 complex is an assembly of CRISPR RNAs and Cas9 endonuclease. The CRISPR RNAs have two separate molecules: (1) guiding CRISPR RNA (crRNA) [95], and (2) trans-acting CRISPR RNA (tracrRNA) [96]. The crRNA recognizes its target DNA sequence and tracrRNA recruits Cas9 to the target sites. Cas9 induces double-strand breaks (DSBs) through its two distinct nuclease domains: an HNH-like nuclease domain that cleaves the gRNA-targeted strand, and a RuvC-like nuclease domain to cut the non-target strand [97]. To simplify the system, the two separate RNAs can be fused into a single RNA chimera, referred to as single-guide RNA (sgRNA), which is able to achieve a similar RNA-guided DNA targeting function [11,98]. SpCas9, a 1368 amino-acid protein from the Class II, Type II category, induces a blunt DSB at the target sequence positioned next to 5′-NGG (N represents any nucleotide) protospacer adjacent motif (PAM) [10]. To increase the availability of target sites and the target specificity, multiple SpCas9 variants have been generated [99], one of which is xCas9 which recognizes 5′-NG, 5′-GAA, and 5′-GAT PAM sequences [100]. To modify the Cas9 enzymatic property, a single point mutation, either D10A or H840A, was introduced into Cas9 to turn it into a nickase (nCas9) that can only cleave single-stranded target sites recognized by gRNA. In contrast to the regular SpCas9 that uses a single sgRNA to produce a DSB, nCas9 requires two sgRNAs to achieve a DSB, thus enhancing the targeting specificity. Another Cas9 derivative is “dead” Cas9 (dCas9) which has a D10A and H840A double mutation. As a result, both HNH- and RuvC-nuclease domains are disabled. This dCas9 binds to its gRNA target site as normal, but not induces breakage in either DNA strand. Thus, the dCas9 can be fused to transcriptional activators or suppressors to regulate gene transcription [101,102].
In addition to SpCas9, a battery of Cas9 orthologues from other bacterial species have been discovered. These orthologues have different PAM sequences that provide a greater choice of target sites across the human genome [103]. Among them are the Cas9 from Streptococcus thermophiles, Neisseria meningitidis, Staphylococcus aureus and Campylobacter jejuni. These different variants of Cas9 offer expanded genome targeting capabilities, improved specificity, and biochemical properties [103,104].
Cas12 and Cas13 from Type V and Type VI categories, respectively, are two latest additions to the gene editing toolbox. Cas12a nucleases from Acidaminococcus spp. (AsCas12a) and Lachnospiraceae spp. (LbCas12a) recognize DNA target sequences with complementarity to the crRNA spacer positioned next to a 3′ PAM [105], and generate a staggered DNA double-strand break by a RuvC domain and a putative nuclease (Nuc) domain. This unique feature makes it ideal for multiplexed genome editing [106]. Cas13 nucleases solely target single-stranded RNA without altering the DNA, which can be selected for transcriptomic manipulation [107].

4.2. Base Editors

Base editing uses a different design to achieve genome modification by the direct generation of precise point mutations. DNA base editors comprise a catalytically impaired Cas nuclease fused with a base modification enzyme that induces base alteration on a single strand of DNA. Then the cellular DNA repair machinery intervenes to repair the mismatch on the complementary strand to complete the intended base conversion [108]. There are two classes of DNA base editors that have been reported: (1) cytosine base editors that convert a C•G base pair into a T•A base pair, and (2) adenine base editors that convert an A•T base pair into a G•C base pair [109].

4.3. Prime Editors

Prime editing enables all types of DNA substitutions, small insertions, and small deletions to be created at targeted sites in the genome [110]. It differs from the CRISPR/Cas9 editing and base editing systems in the following aspects: (1) prime editing does not require DSBs; (2) the prime editor protein is a fusion of nCas9 and a reverse transcriptase; and (3) prime editing guide RNA (pegRNA) contains a specific genome-targeting sequence and a programmable reverse transcriptase template to introduce the desired edit [111]. Compared to Cas nucleases and base editors, prime editors offer several advantages: (1) High precision. As the desired edit is programmed within the pegRNA template, and the targeted DNA is only nicked on one strand, the achieved base change is very specific and rarely generates insertion or deletion mutations. (2) Flexibility. Prime editing can change bases relatively far from the PAM site and is thus less restricted by PAM site availability [112]. (3) Lower cellular replication dependence. Prime editors do not rely on homologous recombination machinery to introduce the desired edit and are therefore effective in any phase of the cell cycle [111].

5. CRISPR-Based CFTR Gene Editing

5.1. In Vitro Correction of CFTR Mutations

Schwank and colleagues first reported using the CRISPR/Cas9 editing system to correct the CFTR locus in cultured intestinal stem cells of F508del-homozygous CF patients [113]. The gene-corrected stem cells were able to develop into organoids that functionally responded to forskolin by volume swelling. Firth et al. derived induced pluripotent stem cells (iPSCs) from skin fibroblast cells from homozygous F508del CF patients, and performed CRISPR/Cas9 gene correction in CFTR. The corrected iPSCs were able to differentiate into mature airway epithelial cells and showed restoration of CFTR-specific chloride transport function [114]. In addition to the prevalent F508del mutation, premature stop codon mutations, such as c.1679 + 1634A > G (1811 + 1.6 kbA > G) and c.3718-2477C > T (3849 + 10 kbC > T), or c.3140-26A > G (3272-26A > G), were corrected by CRISPR/Cas9 editing [115]. Furthermore, the targeted insertion of a full-length CFTR cDNA into the CF locus was accomplished in airway stem cells via CRISPR/Cas9 editing [116]. Similarly, Hu and his team have successfully integrated a full-length CFTR cDNA into GGTA1, a safe harbor gene in cultured pig cells [117]. These data demonstrate the possibility of pan-mutation correction with one gene editing vector.
Using adenine base editors, Krishnamurthy et al. successfully corrected premature stop codon mutations [118]. Moreover, prime editors were recently tested in the functional correction of CF organoids [119]. These seminal experiments clearly demonstrate that different gene editing tools are capable of correcting the basic CF defect in vitro.

5.2. Creation of CF Models

When discussing gene editing for CF, it is hard not to mention the many CF models created using this technology. Because of the simplicity in design and high efficiency in genome manipulation, the gene editing approach has become mainstream in the production of new CF models from cell lines to live animals.
Multiple CF cell lines have been generated using the Cas9 nucleases, including Calu-3 CF cells [120], F508del-CF HL-60 cells [121], F508del-CF T84 cells [122], 16HBE14o- CF cells with F508del, G542X or W1282X mutations [123], CFTR-/- IPEC-J2 porcine cell line [117], and G551D-CF Caco-2 cells [124]. As compared to existing CF cell lines that were immortalized from patients’ tissues, the gene-edited cell lines have their own isogenic controls, thus providing a paired cell culture system for CF basic research and drug selection.
Based on the superior genome-targeting and gene-editing efficiency, direct pronuclear injection or nucleofection of gene editors into fertilized eggs at the single-cell stage (zygote), followed by after-birth screening, is established as a protocol to produce animal models. Using the strategy, multiple CF animal models have been generated, including G542X mice and rats [125,126,127], F508del rabbits [128], and G551D ferrets [129]. Furthermore, by combining CRISPR/Cas9 gene editing and somatic cell nuclear transfer technologies, F508del and G542X CF lambs have been produced. These animal models add valuable selections to the existing CF animal pool to facilitate the study of CF pathogenesis and the development of novel therapies.

5.3. In Vivo Correction of CFTR Mutations

The concept of using different gene editing tools to convey genomic revision of CFTR mutations has been proven by numerous in vitro gene correction cases, as discussed above. Now, the compelling call is to adapt the technology for in vivo CFTR correction as a therapy. As CF lung disease is the major complication in CF patients, early explorations of in vivo gene editing related to CF have been focused on the lung. McCray and his colleagues tested a group of amphiphilic shuttle peptides to facilitate gene editor delivery to mouse lungs and achieved editing of loxP sites in airway epithelia of ROSAmT/mG mice [130]. The editing efficiency was ~13% in the large airways and ~12% in the small airways. The same team also delivered a base editor to Rhesus monkey lungs [131], and achieved editing efficiencies of up to 5.3% in rhesus airway epithelia. In mice, a similar procedure led to the persistence of the gene-edited epithelia for up to 12 months. These exciting early data demonstrate the therapeutic potential and feasibility of using CRISPR-based gene editors for CF lung therapy. It is anticipated that the next step would be applying therapeutic editors to CF animal models to determine if CF lung phenotypes can be rescued.

6. Non-CRISPR-Based CFTR Gene Editing

Although this review is tasked with updating the advances in CRISPR-based gene editing for CF, new developments on the non-CRISPR gene-editing front are worth discussion and comparison. Early non-CRISPR editors include mainly zinc finger nuclease- (ZFN-) based and transcription activator-like effector nuclease- (TALEN-) based. Although currently fewer labs are using these systems for CF gene editing, progress is still being made. Suzuki et al. electroporated ZFN mRNA into CF patient-derived airway basal cells, followed by AAV-6 donor transduction, and achieved efficient functional correction of CFTR [132]. Xia et al. packaged the TALEN system into helper-dependent adenoviral vectors and ~5% targeted gene integration was obtained [133]. In addition to these conventional non-CRISPR editor systems, Egan and her team have adapted a novel peptide nucleic acid (PNA) gene editing system for CFTR gene correction [134,135]. The PNA system was first reported by Rogers et al. to achieve site-specific DNA recognition and to mediate site-directed recombination [136]. Different from the CRISPR-nuclease-triggered endogenous DNA repair, PNAs form a triplex structure with DNA by strand invasion and prompt DNA repair and recombination of short donor DNA [137]. When PNAs and donor DNAs are packed into biodegradable polymer nanoparticles, F508del-CF airway epithelial cells can be targeted and corrected [134]. Notably, using this approach Piotrowski-Daspit and colleagues have achieved in vivo correction of F508del-CFTR mutation in the CF mouse model through systemic delivery [138]. Gene editing has been found in multiple epithelial tissues, including the nasal epithelium, trachea, lung, ileum, colon, and rectum, with a correction rate varying from ~0.1% to ~2%. Impressively, phenotypes such as CF lung inflammation and epithelial chloride transport are shown to be improved.

7. Comparison of CRISPR-Editors and Non-CRISPR PNA Editors for CF Gene Editing

Previous publications have compared the CRISPR editing system with the non-CRISPR TALEN or ZFN system for CF gene correction [135,139]. Here we focus our comparison of the CRISPR system with the new non-CRISPR PNA one. Both systems have been proven feasible in CFTR gene correction in vitro and in vivo, indicating the possibility of rescuing CF defects. However, they are still at their early development stage for in vivo applications, and further optimization and better understanding are clearly needed. First, the rates of editing or correction are still low for a complete rescue of CF. Second, the spectrum of target cells needs a better understanding. Third, the durability of gene editing requires longer observation and inspection. Fourth, long-term safety has not been investigated. Nevertheless, these two approaches are re-igniting the hope of a better CF genetic therapy. Because of their distinct acting mechanisms in genome modification, each approach has its pros and cons, which might be important to consider for future preclinical and clinical trials. First, the CRISPR-based approach needs suitable PAM sites, but the non-CRISPR-based approach does not. Thus, the latter approach may provide better flexibility. Second, the CRISPR-based approach does not have to rely on endogenous DNA repair machinery, but the non-CRISPR-based approach requires this machinery. For example, the base-editing or prime-editing complex carries its own DNA modification enzyme, which may offer a higher editing efficiency. Third, the non-CRISPR editor is small and can fit into nanoparticle carriers. Thus, it can readily pass through tissue barriers to target the epithelial layer. This merit makes systemic delivery possible. Fourth, the non-CRISPR editor does not have high immunogenicity, thus making repetitive administration possible.

8. Barriers to Overcome to Improve In Vivo Gene Editing Efficiency for CF

From the existing data, there is an apparent gap in gene editing efficiency between in vitro and in vivo applications in both CRISPR- and non-CRIPSR-based gene editors. Thus, delivery is a determining factor in the success of translating this new technology into any clinical benefits. As CF predominantly affects epithelium-lined organs, directing CF gene editors to the epithelial cells is believed to be essential for potential CF therapy. There are only two feasible routes to deliver the gene editors: (1) through epithelial lumens, and (2) through the circulation. Each route has its own barriers to overcome.

8.1. Gene Editor Delivery through Epithelial Lumens

For a luminally released gene editor to reach the nucleus of a target epithelial cell, multiple layers of barriers need to be overcome (Figure 1). First, on the top of an epithelium, there usually exists a mucus and/or surface liquid layer that gene editors have to penetrate. Second, an epithelium-lined organ usually evolves a mucociliary clearance mechanism for the purpose of host defense. This mechanism can also act to clear luminally delivered gene editors. Third, epithelial cells are typically polarized, and their subcellular structures, e.g., cytoskeleton, in the apical and basolateral compartments differ, which limits gene editor entry from the apical side. Fourth, epithelial stem and progenitor cells need to be targeted for any lasting gene editing. These cells sometimes are anatomically located underneath other epithelial cells and are hard to reach without losing the overlaying cells.

8.2. Gene Editor Delivery through the Circulatio

Circulation delivery has the potential of targeting every type of cell across the entire body. However, several barriers have to be overcome to achieve any high-level gene editing (Figure 2). First, intravenously released gene editors have to permeate the endothelial layer and escape the circulation to enter the interstitial space. Second, the escaped gene editors need to diffuse through the interstitial space and meshwork to reach the epithelial basement membrane. Third, penetration of the epithelial basement membrane will allow the gene editors to target the epithelial cells from the basolateral side.
Altogether, different routes of delivery dictate the required features in gene editor and vehicle designs. Generally speaking, a smaller size of a gene editor complex favors its penetration and diffusion to reach the target cells. Epithelial and endothelial junctions can be temporally manipulated to facilitate the escape of the editor complex. Vehicles can be engineered to enhance their penetrating and targeting capacity, such as the surface expression of shuttle peptides or receptors/ligands.

9. Prospects for a Cure in CF

Genetic therapy for CF has been pursued for decades [135,139]. The early strategy was focused on gene addition by supplementing the diseased cells with a functional copy of CFTR [140,141,142]. Multiple vectors have been developed and tested pre-clinically or clinically in the lungs, including adenoviral vectors [143,144], adeno-associated viral vectors [145,146,147,148], lentiviral vectors [149,150,151,152], and non-viral vectors [153]. Unfortunately, clinical improvement of lung functions has never been achieved. Retrospective reflection often points to multiple issues, including vector transduction efficiency, stem and progenitor cell targetability, and duration of the transgene expression [141,154]. However, a key and largely overlooked issue is that we do not know whether restoration of CFTR function in epithelial cells alone is sufficient to rescue CF clinical diseases. Nevertheless, there are several outstanding merits of this strategy. First, gene addition by providing a full-length CFTR cDNA can treat all CF mutations at once. Second, after decades of improvements in vector design and delivery procedure, impactful restoration of chloride transport defect across CF epithelia in vivo can be achieved. Third, persistent gene transfer is doable with integrating vectors. These merits maintain gene addition as a viable platform for CF therapy. Moreover, the developed vector systems can be repurposed to deliver gene editors into target cells and organs [155].
The emerging CRISPR-based gene editing technology and the new non-CRISPR PNA-based gene editing technology offer precise and versatile options for genome manipulation, which enable the correction of CF mutations in their natural locations on the chromosome. After this correction, CFTR gene function and regulation should be restored to normal levels. These strategies hold the great promise of achieving an ultimate cure for CF. With the principle proven, translation of the technology into a CF therapy boils down to the following issues: (1) what vehicle to use to best deliver the gene editors, (2) how to effectively target the widespread diseased organ systems, (3) when to perform gene editing to achieve maximal clinical benefits, and (4) whether gene editing is durable and safe. Addressing these issues will bring a true cure for CF within reach.

Funding

This work was supported by a grant from the National Heart, Lung and Blood Institute of the National Institutes of Health (HL150370 to GW).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The author would like to thank Scott Jennings for his comments on and proofreading of this manuscript.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Ishino, Y.; Shinagawa, H.; Makino, K.; Amemura, M.; Nakata, A. Nucleotide sequence of the iap gene, responsible for alkaline phosphatase isozyme conversion in Escherichia coli, and identification of the gene product. J. Bacteriol. 1987, 169, 5429–5433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Mojica, F.J.; Diez-Villasenor, C.; Garcia-Martinez, J.; Soria, E. Intervening sequences of regularly spaced prokaryotic repeats derive from foreign genetic elements. J. Mol. Evol. 2005, 60, 174–182. [Google Scholar] [CrossRef] [PubMed]
  3. Masepohl, B.; Gorlitz, K.; Bohme, H. Long tandemly repeated repetitive (LTRR) sequences in the filamentous cyanobacterium Anabaena sp. PCC 7120. Biochim. Biophys. Acta 1996, 1307, 26–30. [Google Scholar] [CrossRef] [PubMed]
  4. Mojica, F.J.; Diez-Villasenor, C.; Soria, E.; Juez, G. Biological significance of a family of regularly spaced repeats in the genomes of Archaea, Bacteria and mitochondria. Mol. Microbiol. 2000, 36, 244–246. [Google Scholar] [CrossRef]
  5. Jansen, R.; Embden, J.D.; Gaastra, W.; Schouls, L.M. Identification of genes that are associated with DNA repeats in prokaryotes. Mol. Microbiol. 2002, 43, 1565–1575. [Google Scholar] [CrossRef] [PubMed]
  6. Gostimskaya, I. CRISPR-Cas9: A History of Its Discovery and Ethical Considerations of Its Use in Genome Editing. Biochem. (Mosc.) 2022, 87, 777–788. [Google Scholar] [CrossRef] [PubMed]
  7. Barrangou, R.; Fremaux, C.; Deveau, H.; Richards, M.; Boyaval, P.; Moineau, S.; Romero, D.A.; Horvath, P. CRISPR provides acquired resistance against viruses in prokaryotes. Science 2007, 315, 1709–1712. [Google Scholar] [CrossRef]
  8. Garneau, J.E.; Dupuis, M.E.; Villion, M.; Romero, D.A.; Barrangou, R.; Boyaval, P.; Fremaux, C.; Horvath, P.; Magadan, A.H.; Moineau, S. The CRISPR/Cas bacterial immune system cleaves bacteriophage and plasmid DNA. Nature 2010, 468, 67–71. [Google Scholar] [CrossRef]
  9. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef] [Green Version]
  10. Jiang, F.; Doudna, J.A. CRISPR-Cas9 Structures and Mechanisms. Annu. Rev. Biophys. 2017, 46, 505–529. [Google Scholar] [CrossRef] [Green Version]
  11. 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]
  12. Cong, L.; Ran, F.A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P.D.; Wu, X.; Jiang, W.; Marraffini, L.A.; et al. Multiplex genome engineering using CRISPR/Cas systems. Science 2013, 339, 819–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Mali, P.; Yang, L.; Esvelt, K.M.; Aach, J.; Guell, M.; DiCarlo, J.E.; Norville, J.E.; Church, G.M. RNA-guided human genome engineering via Cas9. Science 2013, 339, 823–826. [Google Scholar] [CrossRef] [Green Version]
  14. Frangoul, H.; Altshuler, D.; Cappellini, M.D.; Chen, Y.S.; Domm, J.; Eustace, B.K.; Foell, J.; de la Fuente, J.; Grupp, S.; Handgretinger, R.; et al. CRISPR-Cas9 Gene Editing for Sickle Cell Disease and beta-Thalassemia. N. Engl. J. Med. 2021, 384, 252–260. [Google Scholar] [CrossRef] [PubMed]
  15. Welsh, M.J.; Boat, T.F.; Tsui, L.C.; Beaudet, A.L. Cystic Fibrosis. In The Metabolic Basis of Inherited Disease; Scriver, C.R., Beaudet, A.L., Sly, W.S., Valle, D., Eds.; McGraw-Hill Inc.: New York, NY, USA, 1995; pp. 3799–3876. [Google Scholar]
  16. Shteinberg, M.; Haq, I.J.; Polineni, D.; Davies, J.C. Cystic fibrosis. Lancet 2021, 397, 2195–2211. [Google Scholar] [CrossRef] [PubMed]
  17. Porteous, D.J.; Dorin, J.R. Cystic fibrosis. 3. Cloning the cystic fibrosis gene: Implications for diagnosis and treatment. Thorax 1991, 46, 46–55. [Google Scholar] [CrossRef] [Green Version]
  18. Rommens, J.M.; Iannuzzi, M.C.; Kerem, B.; Drumm, M.L.; Melmer, G.; Dean, M.; Rozmahel, R.; Cole, J.L.; Kennedy, D.; Hidaka, N.; et al. Identification of the cystic fibrosis gene: Chromosome walking and jumping. Science 1989, 245, 1059–1065. [Google Scholar] [CrossRef]
  19. Anderson, M.P.; Gregory, R.J.; Thompson, S.; Souza, D.W.; Paul, S.; Mulligan, R.C.; Smith, A.E.; Welsh, M.J. Demonstration that CFTR is a chloride channel by alteration of its anion selectivity. Science 1991, 253, 202–205. [Google Scholar] [CrossRef] [Green Version]
  20. Kerem, B.; Rommens, J.M.; Buchanan, J.A.; Markiewicz, D.; Cox, T.K.; Chakravarti, A.; Buchwald, M.; Tsui, L.C. Identification of the cystic fibrosis gene: Genetic analysis. Science 1989, 245, 1073–1080. [Google Scholar] [CrossRef] [Green Version]
  21. Cutting, G.R. Cystic fibrosis genetics: From molecular understanding to clinical application. Nat. Rev. Genet. 2015, 16, 45–56. [Google Scholar] [CrossRef] [Green Version]
  22. Sosnay, P.R.; Raraigh, K.S.; Gibson, R.L. Molecular Genetics of Cystic Fibrosis Transmembrane Conductance Regulator: Genotype and Phenotype. Pediatr. Clin. N. Am. 2016, 63, 585–598. [Google Scholar] [CrossRef] [PubMed]
  23. Solomon, G.M.; Marshall, S.G.; Ramsey, B.W.; Rowe, S.M. Breakthrough therapies: Cystic fibrosis (CF) potentiators and correctors. Pediatr. Pulmonol. 2015, 50 (Suppl. 40), S3–S13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Veit, G.; Avramescu, R.G.; Chiang, A.N.; Houck, S.A.; Cai, Z.; Peters, K.W.; Hong, J.S.; Pollard, H.B.; Guggino, W.B.; Balch, W.E.; et al. From CFTR biology toward combinatorial pharmacotherapy: Expanded classification of cystic fibrosis mutations. Mol. Biol. Cell. 2016, 27, 424–433. [Google Scholar] [CrossRef] [Green Version]
  25. Foundation, C.F. Cystic Fibrosis Foundation Patient Rigestry: 2021 Annual Data Report; Cystic Fibrosis Foundation: Bethesda, MD, USA, 2021. [Google Scholar]
  26. Davis, P.B.; Drumm, M.; Konstan, M.W. Cystic fibrosis. Am. J. Respir. Crit. Care Med. 1996, 154, 1229–1256. [Google Scholar] [CrossRef] [PubMed]
  27. Stoltz, D.A.; Meyerholz, D.K.; Welsh, M.J. Origins of cystic fibrosis lung disease. N. Engl. J. Med. 2015, 372, 1574–1575. [Google Scholar] [CrossRef] [Green Version]
  28. Boucher, R.C. Cystic fibrosis: A disease of vulnerability to airway surface dehydration. Trends Mol. Med. 2007, 13, 231–240. [Google Scholar] [CrossRef]
  29. Andersen, D.H. Cystic fibrosis of the pancreas and its relation to celiac disease: A clinical and pathological study. Am. J. Dis. Child. 1938, 56, 344–399. [Google Scholar] [CrossRef]
  30. Gibson-Corley, K.N.; Meyerholz, D.K.; Engelhardt, J.F. Pancreatic pathophysiology in cystic fibrosis. J. Pathol. 2016, 238, 311–320. [Google Scholar] [CrossRef] [Green Version]
  31. Moran, A.; Dunitz, J.; Nathan, B.; Saeed, A.; Holme, B.; Thomas, W. Cystic fibrosis-related diabetes: Current trends in prevalence, incidence, and mortality. Diabetes Care 2009, 32, 1626–1631. [Google Scholar] [CrossRef] [Green Version]
  32. Sathe, M.; Houwen, R. Meconium ileus in Cystic Fibrosis. J. Cyst. Fibros. 2017, 16 (Suppl. 2), S32–S39. [Google Scholar] [CrossRef] [Green Version]
  33. Sankararaman, S.; Schindler, T.; Sferra, T.J. Management of Exocrine Pancreatic Insufficiency in Children. Nutr. Clin. Pract. 2019, 34 (Suppl. 1), S27–S42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Houwen, R.H.; van der Doef, H.P.; Sermet, I.; Munck, A.; Hauser, B.; Walkowiak, J.; Robberecht, E.; Colombo, C.; Sinaasappel, M.; Wilschanski, M.; et al. Defining DIOS and constipation in cystic fibrosis with a multicentre study on the incidence, characteristics, and treatment of DIOS. J. Pediatr. Gastroenterol. Nutr. 2010, 50, 38–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Kelly, T.; Buxbaum, J. Gastrointestinal Manifestations of Cystic Fibrosis. Dig. Dis. Sci. 2015, 60, 1903–1913. [Google Scholar] [CrossRef] [PubMed]
  36. Dorsey, J.; Gonska, T. Bacterial overgrowth, dysbiosis, inflammation, and dysmotility in the Cystic Fibrosis intestine. J. Cyst. Fibros. 2017, 16 (Suppl. 2), S14–S23. [Google Scholar] [CrossRef] [Green Version]
  37. Rogers, G.B.; Narkewicz, M.R.; Hoffman, L.R. The CF gastrointestinal microbiome: Structure and clinical impact. Pediatr. Pulmonol. 2016, 51, S35–S44. [Google Scholar] [CrossRef] [Green Version]
  38. Haber, H.P.; Benda, N.; Fitzke, G.; Lang, A.; Langenberg, M.; Riethmuller, J.; Stern, M. Colonic wall thickness measured by ultrasound: Striking differences in patients with cystic fibrosis versus healthy controls. Gut 1997, 40, 406–411. [Google Scholar] [CrossRef] [Green Version]
  39. Taussig, L.M.; Saldino, R.M.; Di Sant’Agnese, P.A. Radiographic abnormalities of the duodenum and small bowel in cystic fibrosis of the pancreas (mucoviscidosis). Radiology 1973, 106, 369–376. [Google Scholar] [CrossRef]
  40. Littlewood, J.M. Update on intestinal strictures. J. R. Soc. Med. 1999, 92 (Suppl. 37), 41–49. [Google Scholar] [CrossRef]
  41. Dodge, J.A.; Macpherson, C. Colonic strictures in cystic fibrosis. J. R. Soc. Med. 1995, 88 (Suppl. 25), 3–8. [Google Scholar]
  42. Phelan, M.S.; Fine, D.R.; Zentler-Munro, P.L.; Hodson, M.E.; Batten, J.C.; Kerr, I.H. Radiographic abnormalities of the duodenum in cystic fibrosis. Clin. Radiol. 1983, 34, 573–577. [Google Scholar] [CrossRef]
  43. Lloyd-Still, J.D.; Beno, D.W.; Kimura, R.M. Cystic fibrosis colonopathy. Curr. Gastroenterol. Rep. 1999, 1, 231–237. [Google Scholar] [CrossRef] [PubMed]
  44. Norkina, O.; Kaur, S.; Ziemer, D.; De Lisle, R.C. Inflammation of the cystic fibrosis mouse small intestine. Am. J. Physiol. Gastrointest. Liver Physiol. 2004, 286, G1032–G1041. [Google Scholar] [CrossRef] [Green Version]
  45. Kobelska-Dubiel, N.; Klincewicz, B.; Cichy, W. Liver disease in cystic fibrosis. Prz. Gastroenterol. 2014, 9, 136–141. [Google Scholar] [CrossRef] [PubMed]
  46. Cohn, J.A.; Strong, T.V.; Picciotto, M.R.; Nairn, A.C.; Collins, F.S.; Fitz, J.G. Localization of the cystic fibrosis transmembrane conductance regulator in human bile duct epithelial cells. Gastroenterology 1993, 105, 1857–1864. [Google Scholar] [CrossRef] [PubMed]
  47. Beuers, U.; Hohenester, S.; de Buy Wenniger, L.J.; Kremer, A.E.; Jansen, P.L.; Elferink, R.P. The biliary HCO3 umbrella: A unifying hypothesis on pathogenetic and therapeutic aspects of fibrosing cholangiopathies. Hepatology 2010, 52, 1489–1496. [Google Scholar] [CrossRef]
  48. Jarzabek, K.; Zbucka, M.; Pepinski, W.; Szamatowicz, J.; Domitrz, J.; Janica, J.; Wolczynski, S.; Szamatowicz, M. Cystic fibrosis as a cause of infertility. Reprod. Biol. 2004, 4, 119–129. [Google Scholar]
  49. Phillipson, G. Cystic fibrosis and reproduction. Reprod. Fertil. Dev. 1998, 10, 113–119. [Google Scholar] [CrossRef]
  50. Tousson, A.; Van Tine, B.A.; Naren, A.P.; Shaw, G.M.; Schwiebert, L.M. Characterization of CFTR expression and chloride channel activity in human endothelia. Am. J. Physiol. 1998, 275, C1555–C1564. [Google Scholar] [CrossRef] [Green Version]
  51. Lamhonwah, A.M.; Bear, C.E.; Huan, L.J.; Kim Chiaw, P.; Ackerley, C.A.; Tein, I. Cystic fibrosis transmembrane conductance regulator in human muscle: Dysfunction causes abnormal metabolic recovery in exercise. Ann. Neurol. 2010, 67, 802–808. [Google Scholar] [CrossRef]
  52. Robert, R.; Norez, C.; Becq, F. Disruption of CFTR chloride channel alters mechanical properties and cAMP-dependent Cl- transport of mouse aortic smooth muscle cells. J. Physiol. 2005, 568, 483–495. [Google Scholar] [CrossRef]
  53. Risse, P.A.; Kachmar, L.; Matusovsky, O.S.; Novali, M.; Gil, F.R.; Javeshghani, S.; Keary, R.; Haston, C.K.; Michoud, M.C.; Martin, J.G.; et al. Ileal smooth muscle dysfunction and remodeling in cystic fibrosis. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 303, G1–G8. [Google Scholar] [CrossRef] [Green Version]
  54. Hume, J.R.; Hart, P.; Levesque, P.C.; Collier, M.L.; Geary, Y.; Warth, J.; Chapman, T.; Horowitz, B. Molecular physiology of CFTR Cl- channels in heart. Jpn. J. Physiol. 1994, 44 (Suppl. 2), S177–S182. [Google Scholar] [PubMed]
  55. Sterling, K.M., Jr.; Shah, S.; Kim, R.J.; Johnston, N.I.; Salikhova, A.Y.; Abraham, E.H. Cystic fibrosis transmembrane conductance regulator in human and mouse red blood cell membranes and its interaction with ecto-apyrase. J. Cell. Biochem. 2004, 91, 1174–1182. [Google Scholar] [CrossRef] [PubMed]
  56. Mattoscio, D.; Evangelista, V.; De Cristofaro, R.; Recchiuti, A.; Pandolfi, A.; Di Silvestre, S.; Manarini, S.; Martelli, N.; Rocca, B.; Petrucci, G.; et al. Cystic fibrosis transmembrane conductance regulator (CFTR) expression in human platelets: Impact on mediators and mechanisms of the inflammatory response. FASEB J. 2010, 24, 3970–3980. [Google Scholar] [CrossRef] [PubMed]
  57. Reznikov, L.R.; Dong, Q.; Chen, J.H.; Moninger, T.O.; Park, J.M.; Zhang, Y.; Du, J.; Hildebrand, M.S.; Smith, R.J.; Randak, C.O.; et al. CFTR-deficient pigs display peripheral nervous system defects at birth. Proc. Natl. Acad. Sci. USA 2013, 110, 3083–3088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Gong, X.D.; Li, J.C.; Cheung, K.H.; Leung, G.P.; Chew, S.B.; Wong, P.Y. Expression of the cystic fibrosis transmembrane conductance regulator in rat spermatids: Implication for the site of action of antispermatogenic agents. Mol. Hum. Reprod. 2001, 7, 705–713. [Google Scholar] [CrossRef] [Green Version]
  59. Painter, R.G.; Valentine, V.G.; Lanson, N.A., Jr.; Leidal, K.; Zhang, Q.; Lombard, G.; Thompson, C.; Viswanathan, A.; Nauseef, W.M.; Wang, G. CFTR Expression in human neutrophils and the phagolysosomal chlorination defect in cystic fibrosis. Biochemistry 2006, 45, 10260–10269. [Google Scholar] [CrossRef] [Green Version]
  60. Di, A.; Brown, M.E.; Deriy, L.V.; Li, C.; Szeto, F.L.; Chen, Y.; Huang, P.; Tong, J.; Naren, A.P.; Bindokas, V.; et al. CFTR regulates phagosome acidification in macrophages and alters bactericidal activity. Nat. Cell Biol. 2006, 8, 933–944. [Google Scholar] [CrossRef]
  61. Yoshimura, K.; Nakamura, H.; Trapnell, B.C.; Chu, C.S.; Dalemans, W.; Pavirani, A.; Lecocq, J.P.; Crystal, R.G. Expression of the cystic fibrosis transmembrane conductance regulator gene in cells of non-epithelial origin. Nucleic Acids Res. 1991, 19, 5417–5423. [Google Scholar] [CrossRef]
  62. Zhou, Y.; Song, K.; Painter, R.G.; Aiken, M.; Reiser, J.; Stanton, B.A.; Nauseef, W.M.; Wang, G. Cystic fibrosis transmembrane conductance regulator recruitment to phagosomes in neutrophils. J. Innate Immun. 2013, 5, 219–230. [Google Scholar] [CrossRef]
  63. Aiken, M.L.; Painter, R.G.; Zhou, Y.; Wang, G. Chloride transport in functionally active phagosomes isolated from Human neutrophils. Free. Radic. Biol. Med. 2012, 53, 2308–2317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Wang, G. Chloride flux in phagocytes. Immunol. Rev. 2016, 273, 219–231. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, G.; Nauseef, W.M. Salt, chloride, bleach, and innate host defense. J. Leukoc. Biol. 2015, 98, 163–172. [Google Scholar] [CrossRef] [Green Version]
  66. Ng, H.P.; Valentine, V.G.; Wang, G. CFTR targeting during activation of human neutrophils. J. Leukoc. Biol. 2016, 100, 1413–1424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Painter, R.G.; Marrero, L.; Lombard, G.A.; Valentine, V.G.; Nauseef, W.M.; Wang, G. CFTR-mediated halide transport in phagosomes of human neutrophils. J. Leukoc. Biol. 2010, 87, 933–942. [Google Scholar] [CrossRef] [Green Version]
  68. Painter, R.G.; Bonvillain, R.W.; Valentine, V.G.; Lombard, G.A.; LaPlace, S.G.; Nauseef, W.M.; Wang, G. The role of chloride anion and CFTR in killing of Pseudomonas aeruginosa by normal and CF neutrophils. J. Leukoc. Biol. 2008, 83, 1345–1353. [Google Scholar] [CrossRef] [Green Version]
  69. Ng, H.P.; Zhou, Y.; Song, K.; Hodges, C.A.; Drumm, M.L.; Wang, G. Neutrophil-mediated phagocytic host defense defect in myeloid cftr-inactivated mice. PLoS ONE 2014, 9, e106813. [Google Scholar] [CrossRef] [Green Version]
  70. Ng, H.P.; Jennings, S.; Wellems, D.; Sun, F.; Xu, J.; Nauseef, W.M.; Wang, G. Myeloid CFTR loss-of-function causes persistent neutrophilic inflammation in cystic fibrosis. J. Leukoc. Biol. 2020, 108, 1777–1785. [Google Scholar] [CrossRef]
  71. Sajjan, U.; Corey, M.; Humar, A.; Tullis, E.; Cutz, E.; Ackerley, C.; Forstner, J. Immunolocalisation of Burkholderia cepacia in the lungs of cystic fibrosis patients. J. Med. Microbiol. 2001, 50, 535–546. [Google Scholar] [CrossRef] [Green Version]
  72. Hartl, D.; Latzin, P.; Hordijk, P.; Marcos, V.; Rudolph, C.; Woischnik, M.; Krauss-Etschmann, S.; Koller, B.; Reinhardt, D.; Roscher, A.A.; et al. Cleavage of CXCR1 on neutrophils disables bacterial killing in cystic fibrosis lung disease. Nat. Med. 2007, 13, 1423–1430. [Google Scholar] [CrossRef]
  73. Su, X.; Looney, M.R.; Su, H.E.; Lee, J.W.; Song, Y.; Matthay, M.A. Role of CFTR expressed by neutrophils in modulating acute lung inflammation and injury in mice. Inflamm. Res. 2011, 60, 619–632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Witko-Sarsat, V.; Allen, R.C.; Paulais, M.; Nguyen, A.T.; Bessou, G.; Lenoir, G.; Descamps-Latscha, B. Disturbed myeloperoxidase-dependent activity of neutrophils in cystic fibrosis homozygotes and heterozygotes, and its correction by amiloride. J. Immunol. 1996, 157, 2728–2735. [Google Scholar] [CrossRef] [PubMed]
  75. Adib-Conquy, M.; Pedron, T.; Petit-Bertron, A.F.; Tabary, O.; Corvol, H.; Jacquot, J.; Clement, A.; Cavaillon, J.M. Neutrophils in cystic fibrosis display a distinct gene expression pattern. Mol. Med. 2008, 14, 36–44. [Google Scholar] [CrossRef] [PubMed]
  76. Tirouvanziam, R.; Gernez, Y.; Conrad, C.K.; Moss, R.B.; Schrijver, I.; Dunn, C.E.; Davies, Z.A.; Herzenberg, L.A. Profound functional and signaling changes in viable inflammatory neutrophils homing to cystic fibrosis airways. Proc. Natl. Acad. Sci. USA 2008, 105, 4335–4339. [Google Scholar] [CrossRef] [Green Version]
  77. Reeves, E.P.; Williamson, M.; O’Neill, S.J.; Greally, P.; McElvaney, N.G. Nebulized hypertonic saline decreases IL-8 in sputum of patients with cystic fibrosis. Am. J. Respir. Crit. Care Med. 2011, 183, 1517–1523. [Google Scholar] [CrossRef]
  78. Corvol, H.; Fitting, C.; Chadelat, K.; Jacquot, J.; Tabary, O.; Boule, M.; Cavaillon, J.M.; Clement, A. Distinct cytokine production by lung and blood neutrophils from children with cystic fibrosis. Am. J. Physiol. Lung Cell. Mol. Physiol. 2003, 284, L997–L1003. [Google Scholar] [CrossRef] [Green Version]
  79. Moriceau, S.; Lenoir, G.; Witko-Sarsat, V. In cystic fibrosis homozygotes and heterozygotes, neutrophil apoptosis is delayed and modulated by diamide or roscovitine: Evidence for an innate neutrophil disturbance. J. Innate Immun. 2010, 2, 260–266. [Google Scholar] [CrossRef]
  80. Marcos, V.; Zhou, Z.; Yildirim, A.O.; Bohla, A.; Hector, A.; Vitkov, L.; Wiedenbauer, E.M.; Krautgartner, W.D.; Stoiber, W.; Belohradsky, B.H.; et al. CXCR2 mediates NADPH oxidase-independent neutrophil extracellular trap formation in cystic fibrosis airway inflammation. Nat. Med. 2010, 16, 1018–1023. [Google Scholar] [CrossRef]
  81. Kettle, A.J.; Turner, R.; Gangell, C.L.; Harwood, D.T.; Khalilova, I.S.; Chapman, A.L.; Winterbourn, C.C.; Sly, P.D. Oxidation contributes to low glutathione in the airways of children with cystic fibrosis. Eur. Respir. J. 2014, 44, 122–129. [Google Scholar] [CrossRef] [Green Version]
  82. Pohl, K.; Hayes, E.; Keenan, J.; Henry, M.; Meleady, P.; Molloy, K.; Jundi, B.; Bergin, D.A.; McCarthy, C.; McElvaney, O.J.; et al. A neutrophil intrinsic impairment affecting Rab27a and degranulation in cystic fibrosis is corrected by CFTR potentiator therapy. Blood 2014, 24, 999–1009. [Google Scholar] [CrossRef]
  83. Bruscia, E.M.; Zhang, P.X.; Satoh, A.; Caputo, C.; Medzhitov, R.; Shenoy, A.; Egan, M.E.; Krause, D.S. Abnormal trafficking and degradation of TLR4 underlie the elevated inflammatory response in cystic fibrosis. J. Immunol. 2011, 186, 6990–6998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Bruscia, E.M.; Zhang, P.X.; Ferreira, E.; Caputo, C.; Emerson, J.W.; Tuck, D.; Krause, D.S.; Egan, M.E. Macrophages directly contribute to the exaggerated inflammatory response in cystic fibrosis transmembrane conductance regulator−/− mice. Am. J. Respir. Cell Mol. Biol. 2009, 40, 295–304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Hartl, D.; Gaggar, A.; Bruscia, E.; Hector, A.; Marcos, V.; Jung, A.; Greene, C.; McElvaney, G.; Mall, M.; Doring, G. Innate immunity in cystic fibrosis lung disease. J. Cyst. Fibros. 2012, 11, 363–382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Deriy, L.V.; Gomez, E.A.; Zhang, G.; Beacham, D.W.; Hopson, J.A.; Gallan, A.J.; Shevchenko, P.D.; Bindokas, V.P.; Nelson, D.J. Disease-causing mutations in the cystic fibrosis transmembrane conductance regulator determine the functional responses of alveolar macrophages. J. Biol. Chem. 2009, 284, 35926–35938. [Google Scholar] [CrossRef] [Green Version]
  87. Del Porto, P.; Cifani, N.; Guarnieri, S.; Di Domenico, E.G.; Mariggio, M.A.; Spadaro, F.; Guglietta, S.; Anile, M.; Venuta, F.; Quattrucci, S.; et al. Dysfunctional CFTR alters the bactericidal activity of human macrophages against Pseudomonas aeruginosa. PLoS ONE 2011, 6, e19970. [Google Scholar] [CrossRef] [Green Version]
  88. Soleti, R.; Porro, C.; Martinez, M.C. Apoptotic process in cystic fibrosis cells. Apoptosis 2013, 18, 1029–1038. [Google Scholar] [CrossRef] [Green Version]
  89. Turner, M.A.; Baildam, E.; Patel, L.; David, T.J. Joint disorders in cystic fibrosis. J. R. Soc. Med. 1997, 90 (Suppl. 31), 13–20. [Google Scholar] [CrossRef]
  90. Botton, E.; Saraux, A.; Laselve, H.; Jousse, S.; Le Goff, P. Musculoskeletal manifestations in cystic fibrosis. Jt. Bone Spine 2003, 70, 327–335. [Google Scholar] [CrossRef]
  91. Thornton, J.; Rangaraj, S. Disease Modifying Anti-Rheumatic Drugs in People with Cystic Fibrosis-Related Arthritis; Cochrane: London, UK, 2012. [Google Scholar] [CrossRef]
  92. Rush, P.J.; Shore, A.; Coblentz, C.; Wilmot, D.; Corey, M.; Levison, H. The musculoskeletal manifestations of cystic fibrosis. Semin. Arthritis Rheum. 1986, 15, 213–225. [Google Scholar] [CrossRef]
  93. Dixey, J.; Redington, A.N.; Butler, R.C.; Smith, M.J.; Batchelor, J.R.; Woodrow, D.F.; Hodson, M.E.; Batten, J.C.; Brewerton, D.A. The arthropathy of cystic fibrosis. Ann. Rheum. Dis. 1988, 47, 218–223. [Google Scholar] [CrossRef] [Green Version]
  94. Koonin, E.V.; Makarova, K.S. Origins and evolution of CRISPR-Cas systems. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 2019, 374, 20180087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Bolotin, A.; Quinquis, B.; Sorokin, A.; Ehrlich, S.D. Clustered regularly interspaced short palindrome repeats (CRISPRs) have spacers of extrachromosomal origin. Microbiology 2005, 151, 2551–2561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Deltcheva, E.; Chylinski, K.; Sharma, C.M.; Gonzales, K.; Chao, Y.; Pirzada, Z.A.; Eckert, M.R.; Vogel, J.; Charpentier, E. CRISPR RNA maturation by trans-encoded small RNA and host factor RNase III. Nature 2011, 471, 602–607. [Google Scholar] [CrossRef] [Green Version]
  97. Chen, H.; Choi, J.; Bailey, S. Cut site selection by the two nuclease domains of the Cas9 RNA-guided endonuclease. J. Biol. Chem. 2014, 289, 13284–13294. [Google Scholar] [CrossRef] [Green Version]
  98. Gasiunas, G.; Barrangou, R.; Horvath, P.; Siksnys, V. Cas9-crRNA ribonucleoprotein complex mediates specific DNA cleavage for adaptive immunity in bacteria. Proc. Natl. Acad. Sci. USA 2012, 109, E2579–E2586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Kleinstiver, B.P.; Prew, M.S.; Tsai, S.Q.; Topkar, V.V.; Nguyen, N.T.; Zheng, Z.; Gonzales, A.P.; Li, Z.; Peterson, R.T.; Yeh, J.R.; et al. Engineered CRISPR-Cas9 nucleases with altered PAM specificities. Nature 2015, 523, 481–485. [Google Scholar] [CrossRef] [Green Version]
  100. Hu, J.H.; Miller, S.M.; Geurts, M.H.; Tang, W.; Chen, L.; Sun, N.; Zeina, C.M.; Gao, X.; Rees, H.A.; Lin, Z.; et al. Evolved Cas9 variants with broad PAM compatibility and high DNA specificity. Nature 2018, 556, 57–63. [Google Scholar] [CrossRef]
  101. Chen, M.; Qi, L.S. Repurposing CRISPR System for Transcriptional Activation. Adv. Exp. Med. Biol. 2017, 983, 147–157. [Google Scholar] [CrossRef]
  102. Dominguez, A.A.; Lim, W.A.; Qi, L.S. Beyond editing: Repurposing CRISPR-Cas9 for precision genome regulation and interrogation. Nat. Rev. Mol. Cell Biol. 2016, 17, 5–15. [Google Scholar] [CrossRef] [Green Version]
  103. Pickar-Oliver, A.; Gersbach, C.A. The next generation of CRISPR-Cas technologies and applications. Nat. Rev. Mol. Cell Biol. 2019, 20, 490–507. [Google Scholar] [CrossRef]
  104. Karvelis, T.; Gasiunas, G.; Siksnys, V. Harnessing the natural diversity and in vitro evolution of Cas9 to expand the genome editing toolbox. Curr. Opin. Microbiol. 2017, 37, 88–94. [Google Scholar] [CrossRef] [PubMed]
  105. Zetsche, B.; Gootenberg, J.S.; Abudayyeh, O.O.; Slaymaker, I.M.; Makarova, K.S.; Essletzbichler, P.; Volz, S.E.; Joung, J.; van der Oost, J.; Regev, A.; et al. Cpf1 is a single RNA-guided endonuclease of a class 2 CRISPR-Cas system. Cell 2015, 163, 759–771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Yan, F.; Wang, W.; Zhang, J. CRISPR-Cas12 and Cas13: The lesser known siblings of CRISPR-Cas9. Cell Biol. Toxicol. 2019, 35, 489–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Gupta, R.; Ghosh, A.; Chakravarti, R.; Singh, R.; Ravichandiran, V.; Swarnakar, S.; Ghosh, D. Cas13d: A New Molecular Scissor for Transcriptome Engineering. Front. Cell Dev. Biol. 2022, 10, 866800. [Google Scholar] [CrossRef] [PubMed]
  108. Huang, T.P.; Newby, G.A.; Liu, D.R. Precision genome editing using cytosine and adenine base editors in mammalian cells. Nat. Protoc. 2021, 16, 1089–1128. [Google Scholar] [CrossRef]
  109. Rees, H.A.; Liu, D.R. Base editing: Precision chemistry on the genome and transcriptome of living cells. Nat. Rev. Genet. 2018, 19, 770–788. [Google Scholar] [CrossRef]
  110. Chen, P.J.; Liu, D.R. Prime editing for precise and highly versatile genome manipulation. Nat. Rev. Genet. 2023, 24, 161–177. [Google Scholar] [CrossRef]
  111. Doman, J.L.; Sousa, A.A.; Randolph, P.B.; Chen, P.J.; Liu, D.R. Designing and executing prime editing experiments in mammalian cells. Nat. Protoc. 2022, 17, 2431–2468. [Google Scholar] [CrossRef]
  112. 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]
  113. Schwank, G.; Koo, B.K.; Sasselli, V.; Dekkers, J.F.; Heo, I.; Demircan, T.; Sasaki, N.; Boymans, S.; Cuppen, E.; van der Ent, C.K.; et al. Functional repair of CFTR by CRISPR/Cas9 in intestinal stem cell organoids of cystic fibrosis patients. Cell Stem Cell 2013, 13, 653–658. [Google Scholar] [CrossRef] [Green Version]
  114. Firth, A.L.; Menon, T.; Parker, G.S.; Qualls, S.J.; Lewis, B.M.; Ke, E.; Dargitz, C.T.; Wright, R.; Khanna, A.; Gage, F.H.; et al. Functional Gene Correction for Cystic Fibrosis in Lung Epithelial Cells Generated from Patient iPSCs. Cell Rep. 2015, 12, 1385–1390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Sanz, D.J.; Hollywood, J.A.; Scallan, M.F.; Harrison, P.T. Cas9/gRNA targeted excision of cystic fibrosis-causing deep-intronic splicing mutations restores normal splicing of CFTR mRNA. PLoS ONE 2017, 12, e0184009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Vaidyanathan, S.; Baik, R.; Chen, L.; Bravo, D.T.; Suarez, C.J.; Abazari, S.M.; Salahudeen, A.A.; Dudek, A.M.; Teran, C.A.; Davis, T.H.; et al. Targeted replacement of full-length CFTR in human airway stem cells by CRISPR-Cas9 for pan-mutation correction in the endogenous locus. Mol. Ther. 2022, 30, 223–237. [Google Scholar] [CrossRef] [PubMed]
  117. Zhou, Z.P.; Yang, L.L.; Cao, H.; Chen, Z.R.; Zhang, Y.; Wen, X.Y.; Hu, J. In Vitro Validation of a CRISPR-Mediated CFTR Correction Strategy for Preclinical Translation in Pigs. Hum. Gene Ther. 2019, 30, 1448–1458. [Google Scholar] [CrossRef] [PubMed]
  118. Krishnamurthy, S.; Traore, S.; Cooney, A.L.; Brommel, C.M.; Kulhankova, K.; Sinn, P.L.; Newby, G.A.; Liu, D.R.; McCray, P.B. Functional correction of CFTR mutations in human airway epithelial cells using adenine base editors. Nucleic Acids Res. 2021, 49, 10558–10572. [Google Scholar] [CrossRef]
  119. Geurts, M.H.; de Poel, E.; Pleguezuelos-Manzano, C.; Oka, R.; Carrillo, L.; Andersson-Rolf, A.; Boretto, M.; Brunsveld, J.E.; van Boxtel, R.; Beekman, J.M.; et al. Evaluating CRISPR-Based Prime Editing for Cancer Modeling and CFTR Repair in Organoids; Life Science Alliance LLC: Baltimore, MD, USA, 2021. [Google Scholar] [CrossRef]
  120. Bellec, J.; Bacchetta, M.; Losa, D.; Anegon, I.; Chanson, M.; Nguyen, T.H. CFTR inactivation by lentiviral vector-mediated RNA interference and CRISPR-Cas9 genome editing in human airway epithelial cells. Curr. Gene Ther. 2015, 15, 447–459. [Google Scholar] [CrossRef]
  121. Jennings, S.; Ng, H.P.; Wang, G. Establishment of a DeltaF508-CF promyelocytic cell line for cystic fibrosis research and drug screening. J. Cyst. Fibros. 2019, 18, 44–53. [Google Scholar] [CrossRef]
  122. Chung, W.Y.; Song, M.; Park, J.; Namkung, W.; Lee, J.; Kim, H.; Lee, M.G.; Kim, J.Y. Generation of DeltaF508-CFTR T84 cell lines by CRISPR/Cas9-mediated genome editing. Biotechnol. Lett. 2016, 38, 2023–2034. [Google Scholar] [CrossRef]
  123. Valley, H.C.; Bukis, K.M.; Bell, A.; Cheng, Y.; Wong, E.; Jordan, N.J.; Allaire, N.E.; Sivachenko, A.; Liang, F.; Bihler, H.; et al. Isogenic cell models of cystic fibrosis-causing variants in natively expressing pulmonary epithelial cells. J. Cyst. Fibros. 2019, 18, 476–483. [Google Scholar] [CrossRef]
  124. Hao, S.; Roesch, E.A.; Perez, A.; Weiner, R.L.; Henderson, L.C.; Cummings, L.; Consiglio, P.; Pajka, J.; Eisenberg, A.; Yeh, L.; et al. Inactivation of CFTR by CRISPR/Cas9 alters transcriptional regulation of inflammatory pathways and other networks. J. Cyst. Fibros. 2019, 18, e1–e5. [Google Scholar] [CrossRef] [Green Version]
  125. McHugh, D.R.; Steele, M.S.; Valerio, D.M.; Miron, A.; Mann, R.J.; LePage, D.F.; Conlon, R.A.; Cotton, C.U.; Drumm, M.L.; Hodges, C.A. A G542X cystic fibrosis mouse model for examining nonsense mutation directed therapies. PLoS ONE 2018, 13, e0199573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Sharma, J.; Abbott, J.; Klaskala, L.; Zhao, G.; Birket, S.E.; Rowe, S.M. A Novel G542X CFTR Rat Model of Cystic Fibrosis Is Sensitive to Nonsense Mediated Decay. Front. Physiol. 2020, 11, 611294. [Google Scholar] [CrossRef] [PubMed]
  127. McCarron, A.; Cmielewski, P.; Reyne, N.; McIntyre, C.; Finnie, J.; Craig, F.; Rout-Pitt, N.; Delhove, J.; Schjenken, J.E.; Chan, H.Y.; et al. Phenotypic Characterization and Comparison of Cystic Fibrosis Rat Models Generated Using CRISPR/Cas9 Gene Editing. Am. J. Pathol. 2020, 190, 977–993. [Google Scholar] [CrossRef]
  128. Yang, D.; Liang, X.; Pallas, B.; Hoenerhoff, M.; Ren, Z.; Han, R.; Zhang, J.; Chen, Y.E.; Jin, J.P.; Sun, F.; et al. Production of CFTR-DeltaF508 Rabbits. Front. Genet. 2020, 11, 627666. [Google Scholar] [CrossRef] [PubMed]
  129. Yan, Z.; Vorhies, K.; Feng, Z.; Park, S.Y.; Choi, S.H.; Zhang, Y.; Winter, M.; Sun, X.; Engelhardt, J.F. Recombinant Adeno-Associated Virus-Mediated Editing of the G551D Cystic Fibrosis Transmembrane Conductance Regulator Mutation in Ferret Airway Basal Cells. Hum. Gene Ther. 2022, 33, 1023–1036. [Google Scholar] [CrossRef]
  130. Krishnamurthy, S.; Wohlford-Lenane, C.; Kandimalla, S.; Sartre, G.; Meyerholz, D.K.; Theberge, V.; Hallee, S.; Duperre, A.M.; Del’Guidice, T.; Lepetit-Stoffaes, J.P.; et al. Engineered amphiphilic peptides enable delivery of proteins and CRISPR-associated nucleases to airway epithelia. Nat. Commun. 2019, 10, 4906. [Google Scholar] [CrossRef] [Green Version]
  131. Kulhankova, K.; Traore, S.; Cheng, X.; Benk-Fortin, H.; Hallee, S.; Harvey, M.; Roberge, J.; Couture, F.; Gross, T.; Newby, G.; et al. Shuttle Peptide Delivers Base Editor RNPs to Rhesus Monkey Airway Epithelial Cells In Vivo; Research Square Company LLC: Durham, NC, USA, 2023. [Google Scholar] [CrossRef]
  132. Suzuki, S.; Crane, A.M.; Anirudhan, V.; Barilla, C.; Matthias, N.; Randell, S.H.; Rab, A.; Sorscher, E.J.; Kerschner, J.L.; Yin, S.; et al. Highly Efficient Gene Editing of Cystic Fibrosis Patient-Derived Airway Basal Cells Results in Functional CFTR Correction. Mol. Ther. 2020, 28, 1684–1695. [Google Scholar] [CrossRef]
  133. Xia, E.; Zhang, Y.; Cao, H.; Li, J.; Duan, R.; Hu, J. TALEN-Mediated Gene Targeting for Cystic Fibrosis-Gene Therapy. Genes 2019, 10, 39. [Google Scholar] [CrossRef] [Green Version]
  134. McNeer, N.A.; Anandalingam, K.; Fields, R.J.; Caputo, C.; Kopic, S.; Gupta, A.; Quijano, E.; Polikoff, L.; Kong, Y.; Bahal, R.; et al. Nanoparticles that deliver triplex-forming peptide nucleic acid molecules correct F508del CFTR in airway epithelium. Nat. Commun. 2015, 6, 6952. [Google Scholar] [CrossRef] [Green Version]
  135. Egan, M.E. Emerging technologies for cystic fibrosis transmembrane conductance regulator restoration in all people with CF. Pediatr. Pulmonol. 2021, 56 (Suppl. 1), S32–S39. [Google Scholar] [CrossRef]
  136. Rogers, F.A.; Vasquez, K.M.; Egholm, M.; Glazer, P.M. Site-directed recombination via bifunctional PNA-DNA conjugates. Proc. Natl. Acad. Sci. USA 2002, 99, 16695–16700. [Google Scholar] [CrossRef] [Green Version]
  137. 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] [PubMed] [Green Version]
  138. Piotrowski-Daspit, A.S.; Barone, C.; Lin, C.Y.; Deng, Y.; Wu, D.; Binns, T.C.; Xu, E.; Ricciardi, A.S.; Putman, R.; Garrison, A.; et al. In vivo correction of cystic fibrosis mediated by PNA nanoparticles. Sci. Adv. 2022, 8, eabo0522. [Google Scholar] [CrossRef] [PubMed]
  139. Roesch, E.A.; Drumm, M.L. Powerful tools for genetic modification: Advances in gene editing. Pediatr. Pulmonol. 2017, 52, S15–S20. [Google Scholar] [CrossRef] [Green Version]
  140. Flotte, T.R.; Ng, P.; Dylla, D.E.; McCray, P.B., Jr.; Wang, G.; Kolls, J.K.; Hu, J. Viral vector-mediated and cell-based therapies for treatment of cystic fibrosis. Mol. Ther. 2007, 15, 229–241. [Google Scholar] [CrossRef] [PubMed]
  141. Yan, Z.; McCray, P.B., Jr.; Engelhardt, J.F. Advances in gene therapy for cystic fibrosis lung disease. Hum. Mol. Genet. 2019, 28, R88–R94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Vu, A.; McCray, P., Jr. New Directions in Pulmonary Gene Therapy. Hum. Gene Ther. 2020, 31, 921–939. [Google Scholar] [CrossRef]
  143. Welsh, M.J.; Zabner, J.; Graham, S.M.; Smith, A.E.; Moscicki, R.; Wadsworth, S. Adenovirus-mediated gene transfer for cystic fibrosis: Part A. Safety of dose and repeat administration in the nasal epithelium. Part B. Clinical efficacy in the maxillary sinus. Hum. Gene Ther. 1995, 6, 205–218. [Google Scholar] [CrossRef]
  144. Zabner, J.; Couture, L.A.; Gregory, R.J.; Graham, S.M.; Smith, A.E.; Welsh, M.J. Adenovirus-mediated gene transfer transiently corrects the chloride transport defect in nasal epithelia of patients with cystic fibrosis. Cell 1993, 75, 207–216. [Google Scholar] [CrossRef]
  145. Halbert, C.L.; Allen, J.M.; Miller, A.D. Adeno-associated virus type 6 (AAV6) vectors mediate efficient transduction of airway epithelial cells in mouse lungs compared to that of AAV2 vectors. J. Virol. 2001, 75, 6615–6624. [Google Scholar] [CrossRef] [Green Version]
  146. Wagner, J.A.; Nepomuceno, I.B.; Messner, A.H.; Moran, M.L.; Batson, E.P.; Dimiceli, S.; Brown, B.W.; Desch, J.K.; Norbash, A.M.; Conrad, C.K.; et al. A phase II, double-blind, randomized, placebo-controlled clinical trial of tgAAVCF using maxillary sinus delivery in patients with cystic fibrosis with antrostomies. Hum. Gene Ther. 2002, 13, 1349–1359. [Google Scholar] [CrossRef]
  147. Moss, R.B.; Rodman, D.; Spencer, L.T.; Aitken, M.L.; Zeitlin, P.L.; Waltz, D.; Milla, C.; Brody, A.S.; Clancy, J.P.; Ramsey, B.; et al. Repeated adeno-associated virus serotype 2 aerosol-mediated cystic fibrosis transmembrane regulator gene transfer to the lungs of patients with cystic fibrosis: A multicenter, double-blind, placebo-controlled trial. Chest 2004, 125, 509–521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Guggino, W.B.; Benson, J.; Seagrave, J.; Yan, Z.; Engelhardt, J.; Gao, G.; Conlon, T.J.; Cebotaru, L. A Preclinical Study in Rhesus Macaques for Cystic Fibrosis to Assess Gene Transfer and Transduction by AAV1 and AAV5 with a Dual-Luciferase Reporter System. Hum. Gene Ther. Clin. Dev. 2017, 28, 145–156. [Google Scholar] [CrossRef] [PubMed]
  149. Cooney, A.L.; Abou Alaiwa, M.H.; Shah, V.S.; Bouzek, D.C.; Stroik, M.R.; Powers, L.S.; Gansemer, N.D.; Meyerholz, D.K.; Welsh, M.J.; Stoltz, D.A.; et al. Lentiviral-Mediated Phenotypic Correction of Cystic Fibrosis Pigs; American Society for Clinical Investigation (ASCI): Rockville, MD, USA, 2016. [Google Scholar] [CrossRef] [Green Version]
  150. Wang, G.; Slepushkin, V.; Zabner, J.; Keshavjee, S.; Johnston, J.C.; Sauter, S.L.; Jolly, D.J.; Dubensky, T.W., Jr.; Davidson, B.L.; McCray, P.B., Jr. Feline immunodeficiency virus vectors persistently transduce nondividing airway epithelia and correct the cystic fibrosis defect [see comments]. J. Clin. Investig. 1999, 104, R55–R62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Wang, G.; Sinn, P.L.; McCray, P.B., Jr. Development of retroviral vectors for gene transfer to airway epithelia. Curr. Opin. Mol. Ther. 2000, 2, 497–506. [Google Scholar] [PubMed]
  152. Sinn, P.L.; Cooney, A.L.; Oakland, M.; Dylla, D.E.; Wallen, T.J.; Pezzulo, A.A.; Chang, E.H.; McCray, P.B., Jr. Lentiviral vector gene transfer to porcine airways. Mol. Ther. Nucleic Acids 2012, 1, e56. [Google Scholar] [CrossRef] [PubMed]
  153. Alton, E.; 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] [Green Version]
  154. Cooney, A.L.; McCray, P.B., Jr.; Sinn, P.L. Cystic Fibrosis Gene Therapy: Looking Back, Looking Forward. Genes 2018, 9, 538. [Google Scholar] [CrossRef] [Green Version]
  155. Bandara, R.A.; Chen, Z.R.; Hu, J. Potential of helper-dependent Adenoviral vectors in CRISPR-cas9-mediated lung gene therapy. Cell Biosci. 2021, 11, 145. [Google Scholar] [CrossRef]
Figure 1. Gene editor delivery from an epithelial lumen. Potential barriers to overcome to achieve highly efficient gene editor delivery from an epithelial lumen.
Figure 1. Gene editor delivery from an epithelial lumen. Potential barriers to overcome to achieve highly efficient gene editor delivery from an epithelial lumen.
Cells 12 01555 g001
Figure 2. Gene editor delivery via circulation. Potential barriers to overcome to achieve highly efficient gene editor delivery from the circulation.
Figure 2. Gene editor delivery via circulation. Potential barriers to overcome to achieve highly efficient gene editor delivery from the circulation.
Cells 12 01555 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, G. Genome Editing for Cystic Fibrosis. Cells 2023, 12, 1555. https://doi.org/10.3390/cells12121555

AMA Style

Wang G. Genome Editing for Cystic Fibrosis. Cells. 2023; 12(12):1555. https://doi.org/10.3390/cells12121555

Chicago/Turabian Style

Wang, Guoshun. 2023. "Genome Editing for Cystic Fibrosis" Cells 12, no. 12: 1555. https://doi.org/10.3390/cells12121555

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