Next Article in Journal
Primary Humoral Immune Deficiencies: Overlooked Mimickers of Chronic Immune-Mediated Gastrointestinal Diseases in Adults
Next Article in Special Issue
Proteomics and Metabolomics for Cystic Fibrosis Research
Previous Article in Journal
lncRNA_Mdeep: An Alignment-Free Predictor for Distinguishing Long Non-Coding RNAs from Protein-Coding Transcripts by Multimodal Deep Learning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cell-Based Therapeutic Approaches for Cystic Fibrosis

by
Pascal Duchesneau
1,
Thomas K. Waddell
1,2,3 and
Golnaz Karoubi
1,4,5,*
1
Latner Thoracic Surgery Research Laboratories, Toronto General Hospital Research Institute, University Health Network, 101 College St., Toronto, ON M5G 1L7, Canada
2
Institute for Biomaterials and Biomedical Engineering, University of Toronto, 164 College Street, Toronto, ON M5S 3G9, Canada
3
Institute of Medical Science, University of Toronto, 1 King’s College Circle, Toronto, ON M5S 1A8, Canada
4
Department of Laboratory Medicine and Pathobiology, University of Toronto, 1 King’s College Circle, Toronto, ON M5S 1A8, Canada
5
Department of Mechanical and Industrial Engineering, University of Toronto, 5 King’s College Road, Toronto, ON M5S 3G8, Canada
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(15), 5219; https://doi.org/10.3390/ijms21155219
Submission received: 29 June 2020 / Revised: 20 July 2020 / Accepted: 20 July 2020 / Published: 23 July 2020
(This article belongs to the Special Issue Therapeutic Approaches for Cystic Fibrosis 2.0)

Abstract

:
Cystic Fibrosis (CF) is a chronic autosomal recessive disease caused by defects in the cystic fibrosis transmembrane conductance regulator gene (CFTR). Cystic Fibrosis affects multiple organs but progressive remodeling of the airways, mucus accumulation, and chronic inflammation in the lung, result in lung disease as the major cause of morbidity and mortality. While advances in management of CF symptoms have increased the life expectancy of this devastating disease, and there is tremendous excitement about the potential of new agents targeting the CFTR molecule itself, there is still no curative treatment. With the recent advances in the identification of endogenous airway progenitor cells and in directed differentiation of pluripotent cell sources, cell-based therapeutic approaches for CF have become a plausible treatment method with the potential to ultimately cure the disease. In this review, we highlight the current state of cell therapy in the CF field focusing on the relevant autologous and allogeneic cell populations under investigation and the challenges associated with their use. In addition, we present advances in induced pluripotent stem (iPS) cell approaches and emerging new genetic engineering methods, which have the capacity to overcome the current limitations hindering cell therapy approaches.

1. Introduction

Lung diseases are major challenges to the health care sector, and the second leading cause of death in our society. For patients with end-stage lung disease, direct costs are high [1] and lung transplantation has become both a cost-effective treatment approach [2] and often the only life-saving option, despite associated risks of mortality and morbidity due to graft rejection and infection. Cystic Fibrosis (CF), caused by mutations in the gene encoding CF transmembrane conductance regulator (CFTR), is the most common life-limiting, autosomal recessive monogenic disease in Caucasian populations [3]. While the loss of CFTR function affects multiple organs including the lungs, pancreas, liver, and intestine, progressive lung disease and respiratory failure are the major cause of morbidity and mortality for most patients [4]. Recent advances in pharmacological agents such as CFTR correctors and potentiators (reviewed in Gentzsch et al. (2018) and Burgener et al. (2018)) [5,6] and other medical advancements, including lung transplantation, have extended the mean survival of CF patients. However, patients are still faced with reduced quality of life, severe pulmonary complications, and the high costs associated with the lifelong intake of drugs. These limitations and the difficulties of the transplant option continue to drive the search for a more fundamental “cure”.
The most important problem in CF is the defective function of CFTR protein in epithelial cells of the smallest airways. Rather than fixing CFTR in the airway cells, cell replacement therapy would replace them. Mechanistically, the approach could reduce disease impact either via replacement of the defective chloride transport, seen with mutation of the CFTR gene, or lessen the impact of secondary mediators of inflammation. Intact but genetically defective epithelium in the CF airway would be selectively targeted for removal, allowing replacement with progenitor cells with corrected CFTR (Figure 1). This strategy is analogous to hematopoietic stem cell transplantation following cytotoxic chemotherapy, to create “space” in the bone marrow niche. The lung is an ideal organ system for cell-therapy approaches, since minimally invasive access by bronchoscopy, allows us to deliver cells and monitor their persistence directly and potentially their efficacy. Encouragingly, Johnson et al. [7] and others [8] showed that correction of CFTR in only a fraction of cells may be sufficient to restore electrophysiological function and permanently improve clinical outcome. Thus, the concept of “cell replacement therapy” for CF continues to be a scientifically valid and clinically relevant goal.
In this review, we will outline the current state of cell-based therapeutic approaches in the CF field. We will first review the existing pre-clinical animal models of CF and their utility in cell therapy. We will highlight the different cell sources used as vectors in these models and the challenges associated with their use. We will then discuss emerging new ‘designer cells’ from pluripotent sources, molecularly engineered to address some of the current limitations.

2. Etiology and Pathophysiology of CF

Cystic Fibrosis is an autosomal recessive disease caused by mutations in the CFTR gene involved in chloride and bicarbonate transport. CF affects multiple organs such as intestine, pancreas, liver and gallbladder but lung disease is the major cause of morbidity and mortality as a result of mucus accumulation, chronic inflammation, and persistent bacterial infection [9,10,11]. Although CF is caused by mutations in a single gene, over 2000 genetic variants have been identified [12]. Those mutations are classified into six groups according to the synthesis, trafficking, and function of CFTR which include (1) no synthesis, (2) defective processing, (3) defective gating, (4) low conductance, (5) low synthesis, and (6) increased turnover. The most common mutation, a deletion of phenylalanine 508 (ΔF508), accounts for approximately 85% of CF cases and can be classified in multiple groups adding to the complexity. Moreover, patients with the same genetic variation may exhibit different clinical phenotypes attributed to environmental factors and modifier genes [13].
With respect to the epithelium, progressive remodeling of the airways ultimately results in structural damage and impaired lung function and it is unclear whether these changes are related to and initiated by infection/inflammation or are a result of CFTR dysfunction [14,15]. Hyperplasia of goblet and basal cells, squamous metaplasia, increased epithelial height, cell shedding, loss of ciliated epithelial cells, and a disorganization of tight junctions and compound cilia have been reported. In addition, extensive structural changes of the small airway epithelia have also been observed, including epithelial shedding and altered barrier integrity (reviewed in De Rose (2018)) [16].
It is also not clear which CFTR-expressing cells are responsible for CF disease thus making targeted therapeutic approaches challenging. Some studies have suggested that CFTR-dependent submucosal gland secretions have an important role in airway innate immunity [17,18]. More recently, rare FoxI1+ pulmonary ionocytes which are rich in CFTR and account for less than 1% of epithelial cells have been described [19,20]. Ionocytes are believed to have an important role in fluid regulation of airway surfaces and appear to be replenished by basal cells. For significant advancements in cell-based therapies for CF, a better understanding of the most important cell populations to target and ways to remove them will be needed.

3. Animal Models of CF

Animal models are an essential tool to study the pathophysiology of CF and for the development of treatment methodologies. Following cloning of the CF gene in 1989, the first animal model was introduced in mice 3 years later [21,22]. These CFTR-deficient mice quickly developed lethal intestinal obstruction and needed to be fed a liquid diet. To facilitate husbandry and improve survival for lung studies, a gut-corrected model was introduced 2 years later [23]. Despite showing some abnormalities in the lung following bacterial challenge, CFTR-deficient mice did not recapitulate the characteristic lung pathophysiology observed in human patients. This was thought to be the result of a redundant chloride transport channel in murine lungs. Since then, multiple animal models, including conditional CFTR knockout models, have been developed in mice, pigs, zebrafish, rats, rabbits, and sheep [24,25,26,27,28,29].
The pig model showed early signs of airway inflammation, airway remodeling, mucus accumulation, and infection with multiple bacterial species and is 100% lethal without ileostomy performed after birth [9,30]. Thus, a second generation CFTR pig was developed with gut-correction [31]. Similarly, to the CF pigs, CFTR-deficient ferrets also show evidence of lung infections in early life necessitating antibiotic treatment prior to weaning and therefore a gut-corrected second generation was developed [32,33]. Sheep and rabbit CF models are relatively new models and have not yet been fully characterized. Since the sheep has similar lung anatomy to humans it could be a promising model while the rabbit could be useful for fast reproduction and availability of multiple antibodies. However, early results in newborn CFTR−/− sheep showed lethal pancreatic fibrosis, intestinal obstruction, and absence of the vas deferens suggesting that a gut-corrected second generation model may be required [27].
Currently, no single animal model recapitulates all aspects of human CF disease and the majority of cell-based therapeutic studies in the lung have been performed in mice. Larger animal models, such as pig and ferret, are resource-intensive and challenging to raise but may be the most useful preclinical models to study cell-based therapies in a chronic lung infection environment. To our knowledge, no cell therapy has been performed in these models with only few studies focusing on gene therapy. As the field moves towards developing clinically relevant cell-based approaches, large animal models of CF will become useful in assessing cell-based therapies. These models will need to both better recapitulate the human disease and allow for testing of conditioning protocols to remove target epithelium and allow for retention and long-term engraftment of transplanted cells.

4. Overview of Cell Types Used in Cell-Based Therapeutic Approaches for the Lung

The mature lung comprises at least 40 morphologically and functionally distinct cell populations including epithelial, inflammatory, stromal, and endothelial cells [34,35]. Epithelial cells include those forming alveolar units, the Type I and II alveolar epithelial cells, and those lining proximal airways, including ciliated, mucous, Club, basal, and pulmonary neuroendocrine cells (PNEC). Identifying epithelial stem cells in the lung has been difficult due to low cell turnover, but by assessing cell proliferation and using ‘lineage tagging’ techniques [34,35,36,37,38,39], various niches have been identified that contain stem or progenitor cells. Despite the marked progress, the field remains limited in its ability to produce therapeutically applicable cell numbers derived from endogenous stem cells hindering their use in cell-based applications.
The majority of lung regenerative medicine studies have focused on the use of exogenous cell types, predominantly on the use of bone marrow cells (BMC) and in particular the adherent stromal population referred to as mesenchymal stem cells and/or mesenchymal stromal cells (MSC). Engraftment and the significant therapeutic effects of MSC has been shown following various human and experimental lung injuries [40,41,42,43,44,45,46,47,48,49,50]. Although in the majority of cases, therapeutic effects have been observed, there is still considerable debate about the fate of the cells [51,52,53,54], the true level of cell engraftment [42,43,44,45,51,52,53,54], and whether transplanted MSC can truly replace epithelial cells.
As the lung is a complex organ, pluripotent cells would serve as ideal therapeutic units for lung regeneration. There have been significant efforts put forth in generating lung epithelium using embryonic stem (ES) cells and induced pluripotent stem (iPS) cells [55,56,57,58,59,60,61,62,63,64], and recent progress in directed differentiation studies has indicated their potential for use as a cell source for treatment in lung injury models [58,60,65,66,67].

5. Mesenchymal Stromal Cells (MSC)

The bone marrow is the main source for hematopoietic stem cells, harbors endothelial progenitor cells as well as MSC, which represent 0.001–0.01% of the nucleated cells in the marrow [68]. Since their identification in the bone marrow, MSC have been isolated in a variety of other tissues including umbilical cord blood, placental, and adipose tissues [69,70,71]. Data suggesting that these cells can engraft and develop into cell types of other organs including the heart [72], brain [73], liver [74], pancreas [75], skin [76], and lung [77,78,79,80,81,82] has been presented. Most of these observations were based upon co-localization of tissue specific phenotypic markers with sex, transgenically, or dye-marked cells. These reports led to wide interest in BMC therapy for lung injury.
Many groups have since reported much lower lung engraftment of MSC leading to some skepticism in the efficiency of these cells to regenerate epithelium [79,83,84,85,86]. The difficulty in reproducibility of the reported engraftment levels and the possibility of cell fusion has added to the controversy surrounding MSC plasticity. This may have been due to differences in methodologies used, lack of proper controls, and artifacts [86]. Despite concerns regarding lack of standardized methodologies and wide range of variability in observed engraftment potential, the biological effects of BMC are less easily ignored. Therapeutic benefit has been shown following various human and experimental lung injuries such as lung transplantation [87], endotoxin-induced acute lung injury [79,88,89,90,91,92,93], asthma [94,95,96], bronchopulmonary dysplasia [97,98], emphysema [99,100], bleomycin-induced fibrosis [78,97,101,102,103,104,105,106], sepsis [45,107,108,109], and ventilator-induced injury [107,110].
Current thinking suggests that while BMC and MSC are not able to fully transdifferentiate, BMC may promote lung tissue repair through the ability to adopt the expression profiles and functional phenotypes of lung cells even if only temporarily, or via paracrine effects—the ability to secrete soluble growth factors, cytokines, and even organelles which can exert their influence on lung repair. Bioactive factors secreted by MSC are known to mediate immunomodulatory, anti-inflammatory, antioxidant, and anti-apoptotic effects. They also contribute to tissue regeneration, angiogenesis, and clearance of microorganisms. Indeed, the secretory and paracrine effects of MSC have continued to be extensively studied, and it is widely accepted in the MSC field that their secretome is predominantly responsible for the intercellular crosstalk between MSC and targeted cells [111,112,113].
Our lab has contributed to this field with identification of a novel subset of bone marrow cells with particular utility in both mice and human airways [40,41,114,115,116,117]. To increase the therapeutic utility of cell replacement therapy, we have undertaken a number of experiments to optimize cell retention within the lung including a recipient conditioning regimen of naphthalene and busulfan treatment prior to transplantation [41,117]. We have been able to improve BMC retention efficiency by a factor of 10-fold, even in CF CFTR knockout recipient animals. With these technical improvements in delivery regimen, we have achieved substantial increases in expression of CFTR mRNA, and detectable CFTR protein. Importantly, we found that treatment of CFTR−/− mice with CFTR+/+ BMC improved bacterial clearance resulting in greater survival of the CFTR−/− mice treated with BMC [41]. However, despite this progress, proving that the transferred CFTR is functional in vivo has been extremely difficult.
In the context of CF, Wang et al., suggested that MSC were able to differentiate into airway epithelial cells when co-cultured with primary human airway epithelial cells under air–liquid-interface (ALI) conditions. Importantly, the authors found that CFTR-corrected MSC (transduced with a lentiviral vector bearing a wild type CFTR gene) from homozygous ΔF508 CF patients were able to contribute to apical Cl- secretion in response to cAMP agonist stimulation [118]. In subsequent studies by the Conese group, human amniotic MSC were also felt to differentiate into airway epithelial cells when co-cultured with CF immortalized airway epithelial cells [119,120]. Carbone et al. showed that MSC sourced from human amniotic mesenchymal stromal cells co-cultured in ALI with CF immortalized airway epithelial cells were able restore some of the basic defects associated with CF. Specifically, co-cultures had more organized tight junctions with increased expression of occludin and ZO-1 and decreased dextran permeability and resumed chloride transport.
Bonfield and colleagues have shown in a lung infection and inflammation mouse model of CF (Pseudomonas aeruginosa and Staphylococcus aureus), that human MSC decreased the bacterial burden and thereby enhanced the ability of the CF lung to resolve the infection potentially through changes in the in vivo production of the antimicrobial peptide LL-37. They also showed in vitro that supernatant from hMSC derived from both bone marrow and adipose tissue reduced bacterial growth of Pseudomonas aeruginosa, Staphylococcus aureus, and Streptococcus pneumoniae [121]. Further evaluation of the mechanistic action of hMSC showed that in vivo hMSC recruit macrophages known to be important in infection resolution and attenuation of the inflammatory response [122]. In vitro, they showed that hMSC decreased pro-inflammatory cytokine production in LPS-stimulated mouse bone marrow-derived macrophages or human peripheral blood mononuclear cells, and upregulate the expression of PPARγ, which is an important regulator of inflammation in chronic inflammatory diseases such as CF [122,123].
Studies have found that MSC secrete different types of extracellular vesicles (EV) believed to account for much of their therapeutic effects [124,125,126,127]. Additionally, more recent reports have explored the role of released EV from MSC as a potential therapeutic application for controlling inflammation in Cystic Fibrosis [128]. Zulueta and colleagues showed that treatment of IB3-1 CF cell line, (an in vitro human model of CF), with EV derived from human lung MSC under basal and inflammatory conditions (TNFα stimulation) downregulated transcription and protein expression of pro-inflammatory cytokines IL-1β, IL-8, IL-6 and upregulated the mRNA expression of PPARγ (a transcription factor controlling anti-inflammatory and antioxidant mechanisms via NF-kB and HO-1). They also observed reduced NF-kB nuclear translocation and increased HO-1 expression confirming the impairment of the downstream inflammation cascade [128].
Despite recent progress in small animal and in vitro studies, the utility of MSC secretome for treatment of CF remains to be determined and there remains skepticism that MSC will ever function as an equal replacement for respiratory epithelium. Additionally, although there have been numerous preclinical studies on using MSC for cell-based therapy in CF, the vast majority of these have been in small animal models. To determine whether MSC cell or secretome delivery is a viable option for treatment of CF, there needs to be further work in large animal models and in clinical trials. In fact, there are currently only two ongoing phase I clinical trials (NCT02866721 and NCT03058068) investigating MSC therapy in CF. The first, Safety and Tolerability Study of Allogeneic Mesenchymal Stem Cell Infusion in Adults With Cystic Fibrosis (CEASE-CF; NCT02866721), is a prospective, single-center, dose-escalation, open-label interventional study to evaluate the safety and tolerability of allogeneic hMSC in 15 clinically stable subjects with CF age ≥ 18 years. Results from this trial are yet to be published. The second trial, Human Mesenchymal Stem Cells Infusion in Patients With Cystic Fibrosis (HAPI; NCT03058068), focuses on the demonstration of the safety of MSC intravenously administered to 15 adult subjects with Cystic Fibrosis with a secondary objective to explore if MSC can improve the symptoms of cystic fibrosis, including lung function, the rate of pulmonary exacerbation, systemic and local inflammation, and symptom-related quality of life. It is important to note that since its initiation, the latter study has been withdrawn as the principal investigator is no longer at the institution.

6. Induced Progenitor-Like Cells (iPL)

Due to the limitations and skepticism surrounding MSC, as an alternate source of autologous cells, we have been working on a novel cell type which we have developed by careful dissection of events underlying reprogramming during iPS cell generation. Transient reprogramming with transcription factors Oct4, Sox2, Klf4, and c-Myc (OSKM) resulted in an intermediate product of the iPS cell process which we have termed “induced progenitor-like” (iPL) cells [129,130]. We noted that iPL cells are highly proliferative but retain epigenetic “memory” that allows return to their original identity upon withdrawal of reprogramming factors. Specifically, we isolated highly purified populations of Club cells from R26-rtTA/Col1a1:tetO-4F2A double transgenic mice enabling expression of OSKM. We used controlled, transient, exogenous activation of the transcription factors with doxycycline, which causes reprogramming towards iPS cells, but turns off the expression of the OSKM factors prior to reaching the commitment point leading to pluripotency. Throughout this process, we have generated and characterized iPL cell populations derived from the Club cells of the proximal airways [129], alveolar type II epithelial cells (AEC-II) of the distal lung parenchyma [130], and pulmonary endothelial cells (Unpublished data). We found that using Club cells as the starting population, these induced progenitor-like (iPL) cells can be expanded 30-fold yet differentiate normally to ciliated cells with functional expression of CFTR [129]. In vivo, we showed successful retention and incorporation of Club cell-derived iPL cells in airways of CFTR-deficient animals with injected iPL cells giving rise to both Club cells and ciliated cells [129]. A schematic depiction of the generation and utility of Club-cell derived iPL cells is shown in Figure 2.
While we have made significant progress in generating and characterizing a viable source of cells that may be applicable in the clinic in the future, there are several challenges that this technology presents. These include the need for optimization of the partial reprogramming in the human system, as well as the required customization of the approach to any given somatic cell type. For CF patients, an additional step of gene correction will be required. Another limiting factor for autologous cell therapy as a whole, remains the time and effort in obtaining the required cells and the cost associated with the treatment. However, from an immunological perspective, autologous transplantation is ideal and thus it remains a valuable avenue to pursue.

7. Embryonic and Induced Pluripotent Stem Cells (iPS)

Unlike autologous cell sources, human pluripotent stem cells (hPSC) hold enormous promise to serve as the source of unlimited therapeutic cells used to treat injuries and degenerative diseases in future cell-based therapies. There have been significant efforts put forth in generating lung epithelium using pluripotent cell sources [55,56,57,58,59,60,61,62,64,65,66,131,132,133], with evidence for their potential use as a cell source for treatment of lung injury models [57,59,65,66,132,134,135,136,137,138]. Induced pluripotent cells are particularly advantageous to embryonic stem cells, as they are associated with less ethical controversy [139,140] and can be used as an autologous cell source.
Most of the focus in the pluripotent stem cell field in relation to CF, has been on generation of proximal lung epithelial cells to develop patient-specific models of CF [58,60,63] and generation of multi-ciliated cells [61,62]. The protocols presented in these studies follow the basic premise of using the embryological development of the lung as a guide [141] but differ in their use of type, concentration, and time of exposure of chemical stimuli and growth factors. As such, most protocols guide pluripotent stem cells through the developmental stages of definitive endoderm, followed by anteriorization to foregut endoderm, subsequent ventralization to generate NKX2.1+ putative lung progenitor cells, and finally maturation to proximal airway epithelium using air–liquid interface (ALI) culture in which cells are basally exposed by to media and apically to air (Figure 3).
While early attempts using simpler 2D culture protocols showed successful differentiation into lung progenitor populations, mature epithelial cells were not obtained [60], with varying CFTR expression amongst iPS cells derived from CF patients [58]. Subsequent studies [61,131] using lower concentration of BMP4 required during the ventralization phase, also resulted in epithelium lacking multi-ciliated cells but did generate CCSP+ Club cells, MUC5AC+ goblet cells, and some functional CFTR+ cells [61]. This protocol was used to differentiate induced pluripotent stem cells (iPSC) from CF patients carrying a homozygous deletion of F508 in the CFTR gene (resulting in a defective processing of CFTR to the cell membrane) and corrected using clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR associated protein 9 (Cas9) genome editing to target corrective sequences to the endogenous CFTR genomic locus, using a completely excisable piggyBac transposase system [131]. The corrected iPS cells were subsequently differentiated to mature airway epithelial cells and analyzed for CFTR currents by whole cell patch clamp methods. In the CF iPS-derived epithelial cells, approximately 50% of the cells responded to stimulation via exposure to a cocktail of forskolin, genistein, and 3-isobutyl-1-methylxanthine (IBMX) [131].
More recent protocols have incorporated supplemental 3D culture in Matrigel as well as NOTCH inhibition, a maturation cue, during the ALI culture phase [62,63,64,142]. These protocols can achieve multi-ciliated cells with ciliary beating frequency similar to that of primary bronchial epithelial cells. These significant advances in directed differentiation protocols have enabled derivation and use of iPS-derived epithelium from CF patients. In their study, McCauley and colleagues generated airway spheroids from patient-specific CF lines (RC2 202 and RC2 204 lines both with homozygous ΔF508 CFTR mutations) as well as CF corrected iPS lines. They showed forskolin-responsive swelling in normal and not CF patient-derived spheroids and were able to rescue the defect in swelling via genome editing in the CF corrected iPS lines [63].
Although human iPS cells hold enormous promise as the source of unlimited therapeutic cells, there remain challenges in the purity and yield of the cells produced. For utility, standardized protocols are required which produce reproducible numbers of CFTR expressing cells potentially including the newly discovered ionocytes. In addition, prior to clinical application of iPS cells, significant hurdles with respect to their safety and acceptance by the immune system must be overcome.

8. Designer Pluripotent Cells

The generation of designer pluripotent cells, in which iPS cells are molecularly and/or genetically engineered for altered or enhanced function offers the possibility to circumvent their existing limitations. Recent advances in genome editing technologies, particularly the CRISPR/Cas system, have facilitated the targeted integration of functional DNA elements into the human genome, thus, extending their research and therapeutic applications [143]. These include approaches to address both the safety as well as immunogenicity of pluripotent cell sources. The tumorigenicity of hPSC, mainly monitored as teratoma formation after in vivo injection, remains a major challenge that needs to be overcome for the application of hPSC in the clinic. Thus, a variety of approaches have been investigated with the aim to identify and eliminate undifferentiated hPSC including chemical ablation [144], targeting and removal of pluripotent-specific hPSC antigens with cytotoxic antibodies [145], genetic modification of tumor-driving genes [146], insertion of cytotoxic suicide genes [147], and the use of small molecule-based selective elimination [148]. More recently, Liang et al. (2018) described a cell-therapy ‘safety’ solution, termed the safe-cell (SF) suicide system. In this system, they achieve safety by transcriptionally linking a drug-inducible suicide system, Herpes Simplex Virus type 1 thymidine kinase (HSV-TK), to the CDK1 cell division essential loci (essential for a cell to divide or to survive) using CRISPR/Cas9 genome editing; allowing complete control over proliferating cells. Therefore, if required, the dividing cells can be arrested or eliminated by treatment with the drug that induces the suicide system (in this case, ganciclovir (GCV) which activates the HSV-TK suicide system) [149].
In addition to iPS tumorigenicity and the risk of teratoma formation, immune rejection due to expression of human leukocyte antigens (HLA) remains a significant problem. These are encoded by a highly polymorphic set of genes and include HLA class I (HLA-A, HLA-B, and HLA-C) and class II (HLA-DP, HLA-DM, HLA-DO, HLA-DQ, and HLA-DR) molecules [150]. Approaches to generate iPSC stocks isolated from HLA homozygous donors are currently under exploration and it is believed to be possible to cover most HLA haplotypes [151], but the recruitment of HLA homozygous donors that serve an entire population is very difficult and require the generation of a large number of cell lines, extensive validation, and stringent regulatory processes [150]. Genetic manipulation of HLA gene expression for instance via knockout of β2-microglobulin (β2m, major component of HLA class I molecules) resulting in reduced expression of HLA genes has shown to be a promising approach [150,152,153]. Another approach has been to overexpress immunosuppressive genes [152,154]. For example, overexpression of immunosuppressive receptors cytotoxic T-lymphocyte-associated protein 4-immunoglobulin (CTLA4-Ig) and programmed death-ligand 1 (PD-L1) in human ESCs (hESCs) have been shown to prevent allogenic immune rejection [154]. Similarly, Deuse et al. demonstrated that human iPSCs lose their immunogenicity when HLA class I and II genes are inactivated via CRISPR/Cas9 disruption of β2m gene and Ciita (master regulator of HLA class II genes), and CD47 is overexpressed [152].
While suicide gene approaches could eliminate tumor forming cells even after cell transplantation, their efficacy remains controversial. Another concern is the safety of genome editing. Our knowledge of human genomic safe harbors (GSHs) is still insufficient, making it difficult to predict the influence of gene integration on nearby genes and vice versa [143]. Similarly, genetic manipulation of HLA and immunosuppressive genes remain to be fully developed, validated, and likely combined with stringent safety measures prior to implementation in the clinic.

9. Conclusions

Cell-based therapy is a promising approach for CF and would be a less invasive alternative to transplantation, which is limited by organ shortage. Autologous cells could be harvested, gene-corrected in vitro, and transplanted back to the patient’s lungs without the need for immunosuppressive drugs. Mature cells taken from patients could be reprogrammed to iPS to facilitate their expansion and re-differentiated in vitro to proximal and distal epithelium. New technologies such as CRISPR/Cas9 have made gene editing simple and efficient where 30–50% allelic correction led to 20–50% CFTR function restoration in ALI culture [155]. Since CRISPR/Cas9 can have off-target effects resulting in proto-oncogene activity [156,157], the potential tumorigenicity of various gene-corrected cells could be tested in vitro prior to transplantation. For added safety, a fail-safe mechanism, such as herpes simplex virus thymidine kinase, could also be introduced to the cells which could be eliminated by ganciclovir in case of tumorigenic proliferation [149,158].
Many groups, including our own, have investigated ways to improve cell engraftment in the lung by various conditioning regimens to disrupt the epithelium. These treatments include naphthalene, bleomycin, polidocanol, or elastase alone or in combination with busulfan or non-lethal irradiation in animal models and have led to increased cell retention in the lung [41,159,160,161]. As we continue to have a better understanding of cell types involved in the pathogenesis of CF, we can begin to develop preconditioning regiments directly targeting specific cells within the airway epithelium. For instance, standard airway depletion with naphthalene may not be sufficient to deplete submucosal glands and will likely require a harsher insult to expose. One significant challenge that needs to be overcome is the fact that preconditioning regimens are currently in preclinical animal models and not clinically viable options. To achieve clinical translation, it will be necessary to establish clinically viable protocols to ensure adequate cell delivery and engraftment. These will likely require 3D geometric data, mapping the pulmonary vasculature and airways, and computational models that characterize fluid flow and mass transport of the human lung to guide optimization of preconditioning regimens by enabling the selection of process parameters that will minimize the existing limitations of chemical detergents. One possibility is to use ex vivo lung perfusion platform [162] to test and develop these complex therapies.
Early studies using bone marrow MSC faced controversy due to the skepticism that they could truly function as lung epithelium. With the advent of ES and iPS, which can be differentiated in vitro, viable cell sources for use in cell therapy may soon be available [65,129,130,159,163,164]. More studies will need to address the functional integration and long-term regenerative capacity of these cells as well as the safety of engrafted cells. Taken together, cell replacement therapy approaches using pluripotent cell sources in combination with genetic engineering of cells for safety and immunogenicity look to be a promising avenue for a curative treatment (Figure 4).
In conclusion, cell-based therapy for CF is a growing field and for successful future translation to the clinic, several limitations must be addressed. These include (1) development of animal models which better recapitulate the human disease, (2) clinically relevant conditioning regimens to enhance cell retention and engraftment, (3) evaluation of autologous cells, in large animal models and clinical trials (4) refinement of directed differentiation protocols to enhance the yield of proximal airway epithelial cells from pluripotent sources, and (5) assessment of feasibility in use of designer cells in CF models. Nevertheless, cell replacement therapy for CF has huge implications and may significantly increase the quality of life for patients. A strategy in which targeted removal of genetically impaired epithelial cells, followed by repopulation with a gene-corrected cohort, could result in ‘re-designed’ lungs, obviating the need to undergo transplantation.

Author Contributions

P.D., T.K.W., and G.K. wrote and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Funding for studies referred to in this review was provided by Cystic Fibrosis Canada, Canadian Institute of Health Research (CIHR) Emerging Team Grant (RMF-92088), and Astellas Clinical/Basic Research Grant.

Conflicts of Interest

The authors have no financial or non-financial conflicts of interest to declare.

Abbreviations

AEC-II Alveolar type II epithelial cells
ALIAir–liquid-interface
BMC Bone marrow cells
BMP4 Bone Morphogenetic Protein 4
cAMPCyclic adenosine monophosphate
Cas9CRISPR associated protein 9
CCSPClub cell secretory protein
CFBECF bronchial epithelial
CFCystic Fibrosis
CFTRCystic fibrosis transmembrane conductance regulator
CiitaClass II, major histocompatibility complex, transactivator
ClChloride anion
cMycc-Myelocytomatosis oncogene cellular homolog
CRISPRClustered regularly interspaced short palindromic repeats
CTLA4-IgCytotoxic T-lymphocyte-associated protein 4-immunoglobulin
ES Embryonic stem
EVs Extracellular vesicles
FoxI1Forkhead box I1
GCVGanciclovir
GSHsGenomic safe harbors
HLAHuman leukocyte antigens
HO-1Heme oxygenase 1
hPSCsHuman pluripotent stem cells
HSCHematopoietic stem cells
HSV-TK Herpes Simplex Virus type 1 thymidine kinase
IBMX3-isobutyl-1-methylxanthine
IL-1βInterleukin 1 beta
IL-10Interleukin 10
IL-6Interleukin 6
IL-8Interleukin 8
IL-1raInterleukin-1 receptor antagonist
iPLInduced Progenitor-Like Cells
iPSInduced pluripotent stem
KLF4Kruppel Like Factor 4
mRNAMessenger Ribonucleic acid
MSC Mesenchymal stem cells
MUC5ACMucin 5AC, Oligomeric Mucus/Gel-Forming
NF-κBNuclear factor-κB
NKX2-1NK2 Homeobox 1
Oct-4Octamer-binding transcription factor 4
ΔF508Deletion of phenylalanine 508
PNECPulmonary neuroendocrine cells
PPARγPeroxisome proliferator-activated receptor γ
Sox2 SRY (sex determining region Y)-box 2
TNFαTumor necrosis factor alpha
wtCFTRWild-type cystic fibrosis transmembrane conductance regulator
ZO-1Zonula occludens-1

References

  1. López-Campos, J.L.; Tan, W.; Soriano, J.B. Global burden of COPD. Respirology 2016, 21, 14–23. [Google Scholar] [CrossRef] [PubMed]
  2. Vasiliadis, H.-M.; Collet, J.-P.; Penrod, J.R.; Ferraro, P.; Poirier, C. A cost-effectiveness and cost-utility study of lung transplantation. J. Heart Lung Transplant. 2005, 24, 1275–1283. [Google Scholar] [CrossRef] [PubMed]
  3. Cantin, A.M.; Hartl, D.; Konstan, M.W.; Chmiel, J.F. Inflammation in cystic fibrosis lung disease: Pathogenesis and therapy. J. Cyst. Fibros. 2015, 14, 419–430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ratjen, F.; Bell, S.C.; Rowe, S.M.; Goss, C.H.; Quittner, A.L.; Bush, A. Cystic fibrosis. Nat. Rev. Dis. Primers 2015, 1, 15010–15019. [Google Scholar] [CrossRef]
  5. Gentzsch, M.; Mall, M.A. Ion Channel Modulators in Cystic Fibrosis. Chest 2018, 154, 383–393. [Google Scholar] [CrossRef]
  6. Burgener, E.B.; Moss, R.B. Cystic fibrosis transmembrane conductance regulator modulators: Precision medicine in cystic fibrosis. Curr. Opin. Pediatr. 2018, 30, 372–377. [Google Scholar] [CrossRef]
  7. Johnson, L.G.; Olsen, J.C.; Sarkadi, B.; Moore, K.L.; Swanstrom, R.; Boucher, R.C. Efficiency of gene transfer for restoration of normal airway epithelial function in cystic fibrosis. Nat. Genet. 1992, 2, 21–25. [Google Scholar] [CrossRef]
  8. Dorin, J.R.; Farley, R.; Webb, S.; Smith, S.N.; Farini, E.; Delaney, S.J.; Wainwright, B.J.; Alton, E.W.; Porteous, D.J. A demonstration using mouse models that successful gene therapy for cystic fibrosis requires only partial gene correction. Gene Ther. 1996, 3, 797–801. [Google Scholar]
  9. Stoltz, D.A.; Meyerholz, D.K.; Pezzulo, A.A.; Ramachandran, S.; Rogan, M.P.; Davis, G.J.; Hanfland, R.A.; Wohlford-Lenane, C.; Dohrn, C.L.; Bartlett, J.A.; et al. Cystic fibrosis pigs develop lung disease and exhibit defective bacterial eradication at birth. Sci. Transl. Med. 2010, 2, 29ra31. [Google Scholar] [CrossRef] [Green Version]
  10. Moran, O.; Zegarra-Moran, O. On the measurement of the functional properties of the CFTR. J. Cyst. Fibros. 2008, 7, 483–494. [Google Scholar] [CrossRef] [Green Version]
  11. Davies, J.C.; Alton, E.W.F.W.; Bush, A. Cystic fibrosis. BMJ 2007, 335, 1255–1259. [Google Scholar] [CrossRef] [PubMed]
  12. Maiuri, L.; Raia, V.; Kroemer, G. Strategies for the etiological therapy of cystic fibrosis. Cell Death Differ. 2017, 24, 1825–1844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Cutting, G.R. Cystic fibrosis genetics: From molecular understanding to clinical application. Nat. Rev. Genet. 2015, 16, 45–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Cohen-Cymberknoh, M.; Kerem, E.; Ferkol, T.; Elizur, A. Airway inflammation in cystic fibrosis: Molecular mechanisms and clinical implications. Thorax 2013, 68, 1157–1162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. De Rose, V.; Burgel, P.-R.; Gaggar, A.; Greene, C. Airway Inflammatory/Immune Responses in COPD and Cystic Fibrosis. Mediat. Inflamm. 2018, 2018, 7280747-3. [Google Scholar] [CrossRef] [Green Version]
  16. De Rose, V.; Molloy, K.; Gohy, S.; Pilette, C.; Greene, C.M. Airway Epithelium Dysfunction in Cystic Fibrosis and COPD. Mediat. Inflamm. 2018, 2018, 1309746. [Google Scholar] [CrossRef] [Green Version]
  17. Wine, J.J.; Joo, N.S. Submucosal glands and airway defense. Proc. Am. Thorac. Soc. 2004, 1, 47–53. [Google Scholar] [CrossRef] [Green Version]
  18. Dajani, R.; Zhang, Y.; Taft, P.J.; Travis, S.M.; Starner, T.D.; Olsen, A.; Zabner, J.; Welsh, M.J.; Engelhardt, J.F. Lysozyme secretion by submucosal glands protects the airway from bacterial infection. Am. J. Respir. Cell Mol. Biol. 2005, 32, 548–552. [Google Scholar] [CrossRef] [Green Version]
  19. Montoro, D.T.; Haber, A.L.; Biton, M.; Vinarsky, V.; Lin, B.; Birket, S.E.; Yuan, F.; Chen, S.; Leung, H.M.; Villoria, J.; et al. A revised airway epithelial hierarchy includes CFTR-expressing ionocytes. Nature 2018, 560, 319–324. [Google Scholar] [CrossRef]
  20. Plasschaert, L.W.; Žilionis, R.; Choo-Wing, R.; Savova, V.; Knehr, J.; Roma, G.; Klein, A.M.; Jaffe, A.B. A single-cell atlas of the airway epithelium reveals the CFTR-rich pulmonary ionocyte. Nature 2018, 560, 377–381. [Google Scholar] [CrossRef]
  21. Riordan, J.R.; Rommens, J.M.; Kerem, B.; Alon, N.; Rozmahel, R.; Grzelczak, Z.; Zielenski, J.; Lok, S.; Plavsic, N.; Chou, J.L. Identification of the cystic fibrosis gene: Cloning and characterization of complementary DNA. Science 1989, 245, 1066–1073. [Google Scholar] [CrossRef] [PubMed]
  22. Snouwaert, J.N.; Brigman, K.K.; Latour, A.M.; Malouf, N.N.; Boucher, R.C.; Smithies, O.; Koller, B.H. An animal model for cystic fibrosis made by gene targeting. Science 1992, 257, 1083–1088. [Google Scholar] [CrossRef] [Green Version]
  23. Zhou, L.; Dey, C.R.; Wert, S.E.; DuVall, M.D.; Frizzell, R.A.; Whitsett, J.A. Correction of lethal intestinal defect in a mouse model of cystic fibrosis by human CFTR. Science 1994, 266, 1705–1708. [Google Scholar] [CrossRef] [PubMed]
  24. Rogers, C.S.; Hao, Y.; Rokhlina, T.; Samuel, M.; Stoltz, D.A.; Li, Y.; Petroff, E.; Vermeer, D.W.; Kabel, A.C.; Yan, Z.; et al. Production of CFTR-null and CFTR-DeltaF508 heterozygous pigs by adeno-associated virus-mediated gene targeting and somatic cell nuclear transfer. J. Clin. Investig. 2008, 118, 1571–1577. [Google Scholar] [CrossRef] [PubMed]
  25. Navis, A.; Marjoram, L.; Bagnat, M. Cftr controls lumen expansion and function of Kupffer’s vesicle in zebrafish. Development 2013, 140, 1703–1712. [Google Scholar] [CrossRef] [Green Version]
  26. Tuggle, K.L.; Birket, S.E.; Cui, X.; Hong, J.; Warren, J.; Reid, L.; Chambers, A.; Ji, D.; Gamber, K.; Chu, K.K.; et al. Characterization of defects in ion transport and tissue development in cystic fibrosis transmembrane conductance regulator (CFTR)-knockout rats. PLoS ONE 2014, 9, e91253. [Google Scholar] [CrossRef] [Green Version]
  27. Fan, Z.; Perisse, I.V.; Cotton, C.U.; Regouski, M.; Meng, Q.; Domb, C.; Van Wettere, A.J.; Wang, Z.; Harris, A.; White, K.L.; et al. A sheep model of cystic fibrosis generated by CRISPR/Cas9 disruption of the CFTR gene. JCI Insight 2018, 3, 1542. [Google Scholar] [CrossRef]
  28. Hodges, C.A.; Cotton, C.U.; Palmert, M.R.; Drumm, M.L. Generation of a conditional null allele for Cftr in mice. Genesis 2008, 46, 546–552. [Google Scholar] [CrossRef] [Green Version]
  29. Tang, Y.; Yan, Z.; Engelhardt, J.F. Viral Vectors, Animal Models, and Cellular Targets for Gene Therapy of Cystic Fibrosis Lung Disease. Hum. Gene Ther. 2020, 31, 524–537. [Google Scholar] [CrossRef]
  30. Ostedgaard, L.S.; Meyerholz, D.K.; Chen, J.-H.; Pezzulo, A.A.; Karp, P.H.; Rokhlina, T.; Ernst, S.E.; Hanfland, R.A.; Reznikov, L.R.; Ludwig, P.S.; et al. The ΔF508 mutation causes CFTR misprocessing and cystic fibrosis-like disease in pigs. Sci. Transl. Med. 2011, 3, 74ra24. [Google Scholar] [CrossRef] [Green Version]
  31. Stoltz, D.A.; Rokhlina, T.; Ernst, S.E.; Pezzulo, A.A.; Ostedgaard, L.S.; Karp, P.H.; Samuel, M.S.; Reznikov, L.R.; Rector, M.V.; Gansemer, N.D.; et al. Intestinal CFTR expression alleviates meconium ileus in cystic fibrosis pigs. J. Clin. Investig. 2013, 123, 2685–2693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Sun, X.; Sui, H.; Fisher, J.T.; Yan, Z.; Liu, X.; Cho, H.-J.; Joo, N.S.; Zhang, Y.; Zhou, W.; Yi, Y.; et al. Disease phenotype of a ferret CFTR-knockout model of cystic fibrosis. J. Clin. Investig. 2010, 120, 3149–3160. [Google Scholar] [CrossRef] [Green Version]
  33. Sun, X.; Olivier, A.K.; Yi, Y.; Pope, C.E.; Hayden, H.S.; Liang, B.; Sui, H.; Zhou, W.; Hager, K.R.; Zhang, Y.; et al. Gastrointestinal pathology in juvenile and adult CFTR-knockout ferrets. Am. J. Pathol. 2014, 184, 1309–1322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Rawlins, E.L.; Clark, C.P.; Xue, Y.; Hogan, B.L.M. The Id2+ distal tip lung epithelium contains individual multipotent embryonic progenitor cells. Development 2009, 136, 3741–3745. [Google Scholar] [CrossRef] [Green Version]
  35. Rawlins, E.L.; Ostrowski, L.E.; Randell, S.H.; Hogan, B.L.M. Lung development and repair: Contribution of the ciliated lineage. Proc. Natl. Acad. Sci. USA 2007, 104, 410–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Barkauskas, C.E.; Cronce, M.J.; Rackley, C.R.; Bowie, E.J.; Keene, D.R.; Stripp, B.R.; Randell, S.H.; Noble, P.W.; Hogan, B.L.M. Type 2 alveolar cells are stem cells in adult lung. J. Clin. Investig. 2013, 123, 3025–3036. [Google Scholar] [CrossRef] [PubMed]
  37. Barth, P.J.; Koch, S.; Müller, B.; Unterstab, F.; Unterstab, F.; von Wichert, P.; Moll, R. Proliferation and number of Clara cell 10-kDa protein (CC10)-reactive epithelial cells and basal cells in normal, hyperplastic and metaplastic bronchial mucosa. Virchows Arch. 2000, 437, 648–655. [Google Scholar] [CrossRef]
  38. Borthwick, D.W.; Shahbazian, M.; Krantz, Q.T.; Dorin, J.R.; Randell, S.H. Evidence for stem-cell niches in the tracheal epithelium. Am. J. Respir. Cell Mol. Biol. 2001, 24, 662–670. [Google Scholar] [CrossRef] [Green Version]
  39. Giangreco, A.; Shen, H.; Reynolds, S.D.; Stripp, B.R. Molecular phenotype of airway side population cells. Am. J. Physiol. Lung Cell Mol. Physiol. 2004, 286, L624–L630. [Google Scholar] [CrossRef]
  40. Bustos, M.L.; Mura, M.; Marcus, P.; Hwang, D.; Ludkovski, O.; Wong, A.P.; Waddell, T.K. Bone marrow cells expressing clara cell secretory protein increase epithelial repair after ablation of pulmonary clara cells. Mol. Ther. 2013, 21, 1251–1258. [Google Scholar] [CrossRef] [Green Version]
  41. Duchesneau, P.; Besla, R.; Derouet, M.F.; Guo, L.; Karoubi, G.; Silberberg, A.; Wong, A.P.; Waddell, T.K. Partial Restoration of CFTR Function in cftr-Null Mice following Targeted Cell Replacement Therapy. Mol. Ther. 2017, 25, 654–665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Weiss, D.J. Concise review: Current status of stem cells and regenerative medicine in lung biology and diseases. Stem Cells 2014, 32, 16–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Antunes, M.A.; Laffey, J.G.; Pelosi, P.; Rocco, P.R.M. Mesenchymal stem cell trials for pulmonary diseases. J. Cell. Biochem. 2014, 115, 1023–1032. [Google Scholar] [CrossRef] [PubMed]
  44. Li, J.; Zhou, J.; Zhang, D.; Song, Y.; She, J.; Bai, C. Bone marrow-derived mesenchymal stem cells enhance autophagy via PI3K/AKT signalling to reduce the severity of ischaemia/reperfusion-induced lung injury. J. Cell. Mol. Med. 2015, 19, 2341–2351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Ho, M.S.H.; Mei, S.H.J.; Stewart, D.J. The Immunomodulatory and Therapeutic Effects of Mesenchymal Stromal Cells for Acute Lung Injury and Sepsis. J. Cell. Physiol. 2015, 230, 2606–2617. [Google Scholar] [CrossRef]
  46. Monsel, A.; Zhu, Y.-G.; Gennai, S.; Hao, Q.; Hu, S.; Rouby, J.-J.; Rosenzwajg, M.; Matthay, M.A.; Lee, J.W. Therapeutic Effects of Human Mesenchymal Stem Cell-derived Microvesicles in Severe Pneumonia in Mice. Am. J. Respir. Crit. Care Med. 2015, 192, 324–336. [Google Scholar] [CrossRef] [Green Version]
  47. Gupta, N.; Sinha, R.; Krasnodembskaya, A.; Xu, X.; Nizet, V.; Matthay, M.A.; Griffin, J.H. The TLR4-PAR1 Axis Regulates Bone Marrow Mesenchymal Stromal Cell Survival and Therapeutic Capacity in Experimental Bacterial Pneumonia. Stem Cells 2018, 36, 796–806. [Google Scholar] [CrossRef] [Green Version]
  48. Cruz, F.F.; Borg, Z.D.; Goodwin, M.; Sokocevic, D.; Wagner, D.E.; Coffey, A.; Antunes, M.; Robinson, K.L.; Mitsialis, S.A.; Kourembanas, S.; et al. Systemic Administration of Human Bone Marrow-Derived Mesenchymal Stromal Cell Extracellular Vesicles Ameliorates Aspergillus Hyphal Extract-Induced Allergic Airway Inflammation in Immunocompetent Mice. Stem Cells Transl. Med. 2015, 4, 1302–1316. [Google Scholar] [CrossRef] [Green Version]
  49. Kitoko, J.Z.; de Castro, L.L.; Nascimento, A.P.; Abreu, S.C.; Cruz, F.F.; Arantes, A.C.; Xisto, D.G.; Martins, M.A.; Morales, M.M.; Rocco, P.R.M.; et al. Therapeutic administration of bone marrow-derived mesenchymal stromal cells reduces airway inflammation without up-regulating Tregs in experimental asthma. Clin. Exp. Allergy 2018, 48, 205–216. [Google Scholar] [CrossRef]
  50. Kennelly, H.; Mahon, B.P.; English, K. Human mesenchymal stromal cells exert HGF dependent cytoprotective effects in a human relevant pre-clinical model of COPD. Sci. Rep. 2016, 6, 38207–38211. [Google Scholar] [CrossRef]
  51. Eisenberg, L.M.; Eisenberg, C.A. Stem cell plasticity, cell fusion, and transdifferentiation. Birth Defects Res. C Embryo Today 2003, 69, 209–218. [Google Scholar] [CrossRef] [PubMed]
  52. Perán, M.; Marchal, J.A.; Rodríguez-Serrano, F.; Alvarez, P.; Aránega, A. Transdifferentiation: Why and how? Cell Biol. Int. 2011, 35, 373–379. [Google Scholar] [CrossRef] [PubMed]
  53. Sanges, D.; Lluis, F.; Cosma, M.P. Cell-fusion-mediated reprogramming: Pluripotency or transdifferentiation? Implications for regenerative medicine. Adv. Exp. Med. Biol. 2011, 713, 137–159. [Google Scholar] [PubMed]
  54. Moodley, Y.; Manuelpillai, U.; Weiss, D.J. Cellular therapies for lung disease: A distant horizon. Respirology 2011, 16, 223–237. [Google Scholar] [CrossRef]
  55. Rippon, H.J.; Lane, S.; Qin, M.; Ismail, N.-S.; Wilson, M.R.; Takata, M.; Bishop, A.E. Embryonic stem cells as a source of pulmonary epithelium in vitro and in vivo. Proc. Am. Thorac. Soc. 2008, 5, 717–722. [Google Scholar] [CrossRef] [PubMed]
  56. Van Vranken, B.E.; Rippon, H.J.; Samadikuchaksaraei, A.; Trounson, A.O.; Bishop, A.E. The differentiation of distal lung epithelium from embryonic stem cells. Curr. Protoc. Stem Cell Biol. 2007, 2, 1G.1.1–1G.1.22. [Google Scholar]
  57. Huang, S.X.L.; Islam, M.N.; O’Neill, J.; Hu, Z.; Yang, Y.-G.; Chen, Y.-W.; Mumau, M.; Green, M.D.; Vunjak-Novakovic, G.; Bhattacharya, J.; et al. Efficient generation of lung and airway epithelial cells from human pluripotent stem cells. Nat. Biotechnol. 2014, 32, 84–91. [Google Scholar] [CrossRef] [Green Version]
  58. Wong, A.P.; Bear, C.E.; Chin, S.; Pasceri, P.; Thompson, T.O.; Huan, L.-J.; Ratjen, F.; Ellis, J.; Rossant, J. Directed differentiation of human pluripotent stem cells into mature airway epithelia expressing functional CFTR protein. Nat. Biotechnol. 2012, 30, 876–882. [Google Scholar] [CrossRef] [Green Version]
  59. Longmire, T.A.; Ikonomou, L.; Hawkins, F.; Christodoulou, C.; Cao, Y.; Jean, J.C.; Kwok, L.W.; Mou, H.; Rajagopal, J.; Shen, S.S.; et al. Efficient derivation of purified lung and thyroid progenitors from embryonic stem cells. Cell Stem Cell 2012, 10, 398–411. [Google Scholar] [CrossRef] [Green Version]
  60. Mou, H.; Zhao, R.; Sherwood, R.; Ahfeldt, T.; Lapey, A.; Wain, J.; Sicilian, L.; Izvolsky, K.; Musunuru, K.; Cowan, C.; et al. Generation of multipotent lung and airway progenitors from mouse ESCs and patient-specific cystic fibrosis iPSCs. Cell Stem Cell 2012, 10, 385–397. [Google Scholar] [CrossRef] [Green Version]
  61. Firth, A.L.; Dargitz, C.T.; Qualls, S.J.; Menon, T.; Wright, R.; Singer, O.; Gage, F.H.; Khanna, A.; Verma, I.M. Generation of multiciliated cells in functional airway epithelia from human induced pluripotent stem cells. Proc. Natl. Acad. Sci. USA 2014, 111, E1723–E1730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Konishi, S.; Gotoh, S.; Tateishi, K.; Yamamoto, Y.; Korogi, Y.; Nagasaki, T.; Matsumoto, H.; Muro, S.; Hirai, T.; Ito, I.; et al. Directed Induction of Functional Multi-ciliated Cells in Proximal Airway Epithelial Spheroids from Human Pluripotent Stem Cells. Stem Cell Rep. 2016, 6, 18–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. McCauley, K.B.; Hawkins, F.; Serra, M.; Thomas, D.C.; Jacob, A.; Kotton, D.N. Efficient Derivation of Functional Human Airway Epithelium from Pluripotent Stem Cells via Temporal Regulation of Wnt Signaling. Cell Stem Cell 2017, 20, 844–857.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. McCauley, K.B.; Hawkins, F.; Kotton, D.N. Derivation of Epithelial-Only Airway Organoids from Human Pluripotent Stem Cells. Curr. Protoc. Stem Cell Biol. 2018, 45, e51. [Google Scholar] [CrossRef]
  65. Miller, A.J.; Hill, D.R.; Nagy, M.S.; Aoki, Y.; Dye, B.R.; Chin, A.M.; Huang, S.; Zhu, F.; White, E.S.; Lama, V.; et al. In Vitro Induction and In Vivo Engraftment of Lung Bud Tip Progenitor Cells Derived from Human Pluripotent Stem Cells. Stem Cell Rep. 2018, 10, 101–119. [Google Scholar] [CrossRef] [Green Version]
  66. Litvack, M.L.; Wigle, T.J.; Lee, J.; Wang, J.; Ackerley, C.; Grunebaum, E.; Post, M. Alveolar-like Stem Cell-derived Myb(-) Macrophages Promote Recovery and Survival in Airway Disease. Am. J. Respir. Crit. Care Med. 2016, 193, 1219–1229. [Google Scholar] [CrossRef]
  67. Wang, Y.-C.; Nakagawa, M.; Garitaonandia, I.; Slavin, I.; Altun, G.; Lacharite, R.M.; Nazor, K.L.; Tran, H.T.; Lynch, C.L.; Leonardo, T.R.; et al. Specific lectin biomarkers for isolation of human pluripotent stem cells identified through array-based glycomic analysis. Cell Res. 2011, 21, 1551–1563. [Google Scholar] [CrossRef] [Green Version]
  68. Salven, P.; Mustjoki, S.; Alitalo, R.; Alitalo, K.; Rafii, S. VEGFR-3 and CD133 identify a population of CD34+ lymphatic/vascular endothelial precursor cells. Blood 2003, 101, 168–172. [Google Scholar] [CrossRef] [Green Version]
  69. Zuk, P.A.; Zhu, M.; Ashjian, P.; De Ugarte, D.A.; Huang, J.I.; Mizuno, H.; Alfonso, Z.C.; Fraser, J.K.; Benhaim, P.; Hedrick, M.H. Human adipose tissue is a source of multipotent stem cells. Mol. Biol. Cell 2002, 13, 4279–4295. [Google Scholar] [CrossRef]
  70. Lee, M.W.; Choi, J.; Yang, M.S.; Moon, Y.J.; Park, J.S.; Kim, H.C.; Kim, Y.J. Mesenchymal stem cells from cryopreserved human umbilical cord blood. Biochem. Biophys. Res. Commun. 2004, 320, 273–278. [Google Scholar] [CrossRef]
  71. Anker, P.S.; Scherjon, S.A.; Kleijburg-van der Keur, C.; de Groot-Swings, G.M.J.S.; Claas, F.H.J.; Fibbe, W.E.; Kanhai, H.H.H. Isolation of mesenchymal stem cells of fetal or maternal origin from human placenta. Stem Cells 2004, 22, 1338–1345. [Google Scholar] [CrossRef] [PubMed]
  72. Kawada, H.; Fujita, J.; Kinjo, K.; Matsuzaki, Y.; Tsuma, M.; Miyatake, H.; Muguruma, Y.; Tsuboi, K.; Itabashi, Y.; Ikeda, Y.; et al. Nonhematopoietic mesenchymal stem cells can be mobilized and differentiate into cardiomyocytes after myocardial infarction. Blood 2004, 104, 3581–3587. [Google Scholar] [CrossRef] [PubMed]
  73. Kopen, G.C.; Prockop, D.J.; Phinney, D.G. Marrow stromal cells migrate throughout forebrain and cerebellum, and they differentiate into astrocytes after injection into neonatal mouse brains. Proc. Natl. Acad. Sci. USA 1999, 96, 10711–10716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Lange, C.; Bassler, P.; Lioznov, M.-V.; Bruns, H.; Kluth, D.; Zander, A.-R.; Fiegel, H.-C. Liver-specific gene expression in mesenchymal stem cells is induced by liver cells. World J. Gastroenterol. 2005, 11, 4497–4504. [Google Scholar] [CrossRef] [PubMed]
  75. Izumida, Y.; Aoki, T.; Yasuda, D.; Koizumi, T.; Suganuma, C.; Saito, K.; Murai, N.; Shimizu, Y.; Hayashi, K.; Odaira, M.; et al. Hepatocyte growth factor is constitutively produced by donor-derived bone marrow cells and promotes regeneration of pancreatic beta-cells. Biochem. Biophys. Res. Commun. 2005, 333, 273–282. [Google Scholar] [CrossRef] [PubMed]
  76. Borue, X.; Lee, S.; Grove, J.; Herzog, E.L.; Harris, R.; Diflo, T.; Glusac, E.; Hyman, K.; Theise, N.D.; Krause, D.S. Bone marrow-derived cells contribute to epithelial engraftment during wound healing. Am. J. Pathol. 2004, 165, 1767–1772. [Google Scholar] [CrossRef] [Green Version]
  77. Krause, D.S.; Theise, N.D.; Collector, M.I.; Henegariu, O.; Hwang, S.; Gardner, R.; Neutzel, S.; Sharkis, S.J. Multi-organ, multi-lineage engraftment by a single bone marrow-derived stem cell. Cell 2001, 105, 369–377. [Google Scholar] [CrossRef] [Green Version]
  78. Kotton, D.N.; Ma, B.Y.; Cardoso, W.V.; Sanderson, E.A.; Summer, R.S.; Williams, M.C.; Fine, A. Bone marrow-derived cells as progenitors of lung alveolar epithelium. Development 2001, 128, 5181–5188. [Google Scholar]
  79. Xu, J.; Woods, C.R.; Mora, A.L.; Joodi, R.; Brigham, K.L.; Iyer, S.; Rojas, M. Prevention of endotoxin-induced systemic response by bone marrow-derived mesenchymal stem cells in mice. Am. J. Physiol. Lung Cell Mol. Physiol. 2007, 293, L131–L141. [Google Scholar] [CrossRef] [Green Version]
  80. Ma, N.; Gai, H.; Mei, J.; Ding, F.-B.; Bao, C.-R.; Nguyen, D.M.; Zhong, H. Bone marrow mesenchymal stem cells can differentiate into type II alveolar epithelial cells in vitro. Cell Biol. Int. 2011, 35, 1261–1266. [Google Scholar] [CrossRef]
  81. Cerrada, A.; de la Torre, P.; Grande, J.; Haller, T.; Flores, A.I.; Pérez-Gil, J. Human decidua-derived mesenchymal stem cells differentiate into functional alveolar type II-like cells that synthesize and secrete pulmonary surfactant complexes. PLoS ONE 2014, 9, e110195. [Google Scholar] [CrossRef] [PubMed]
  82. Li, Y.; Xu, W.; Yan, J.; Xia, Y.; Gu, C.; Ma, Y.; Tao, H. Differentiation of human amniotic fluid-derived mesenchymal stem cells into type II alveolar epithelial cells in vitro. Int. J. Mol. Med. 2014, 33, 1507–1513. [Google Scholar] [CrossRef]
  83. Ortiz, L.A.; Gambelli, F.; McBride, C.; Gaupp, D.; Baddoo, M.; Kaminski, N.; Phinney, D.G. Mesenchymal stem cell engraftment in lung is enhanced in response to bleomycin exposure and ameliorates its fibrotic effects. Proc. Natl. Acad. Sci. USA 2003, 100, 8407–8411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. MacPherson, H.; Keir, P.A.; Edwards, C.J.; Webb, S.; Dorin, J.R. Following damage, the majority of bone marrow-derived airway cells express an epithelial marker. Respir. Res. 2006, 7, 145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Aliotta, J.M.; Keaney, P.; Passero, M.; Dooner, M.S.; Pimentel, J.; Greer, D.; Demers, D.; Foster, B.; Peterson, A.; Dooner, G.; et al. Bone marrow production of lung cells: The impact of G-CSF, cardiotoxin, graded doses of irradiation, and subpopulation phenotype. Exp. Hematol. 2006, 34, 230–241. [Google Scholar] [CrossRef] [Green Version]
  86. Kassmer, S.H.; Krause, D.S. Detection of bone marrow-derived lung epithelial cells. Exp. Hematol. 2010, 38, 564–573. [Google Scholar] [CrossRef] [Green Version]
  87. Kleeberger, W.; Versmold, A.; Rothämel, T.; Glöckner, S.; Bredt, M.; Haverich, A.; Lehmann, U.; Kreipe, H. Increased chimerism of bronchial and alveolar epithelium in human lung allografts undergoing chronic injury. Am. J. Pathol. 2003, 162, 1487–1494. [Google Scholar] [CrossRef] [Green Version]
  88. Gupta, N.; Su, X.; Popov, B.; Lee, J.W.; Serikov, V.; Matthay, M.A. Intrapulmonary delivery of bone marrow-derived mesenchymal stem cells improves survival and attenuates endotoxin-induced acute lung injury in mice. J. Immunol. 2007, 179, 1855–1863. [Google Scholar] [CrossRef] [Green Version]
  89. Iyer, S.S.; Torres-Gonzalez, E.; Neujahr, D.C.; Kwon, M.; Brigham, K.L.; Jones, D.P.; Mora, A.L.; Rojas, M. Effect of bone marrow-derived mesenchymal stem cells on endotoxin-induced oxidation of plasma cysteine and glutathione in mice. Stem Cells Int. 2010, 2010, 868076. [Google Scholar] [CrossRef] [Green Version]
  90. Lee, J.W.; Fang, X.; Gupta, N.; Serikov, V.; Matthay, M.A. Allogeneic human mesenchymal stem cells for treatment of E. coli endotoxin-induced acute lung injury in the ex vivo perfused human lung. Proc. Natl. Acad. Sci. USA 2009, 106, 16357–16362. [Google Scholar] [CrossRef] [Green Version]
  91. Tai, W.-L.; Dong, Z.-X.; Zhang, D.-D.; Wang, D.-H. Therapeutic effect of intravenous bone marrow-derived mesenchymal stem cell transplantation on early-stage LPS-induced acute lung injury in mice. Nan Fang Yi Ke Da Xue Xue Bao 2012, 32, 283–290. [Google Scholar] [PubMed]
  92. Xu, J.; Qu, J.; Cao, L.; Sai, Y.; Chen, C.; He, L.; Yu, L. Mesenchymal stem cell-based angiopoietin-1 gene therapy for acute lung injury induced by lipopolysaccharide in mice. J. Pathol. 2008, 214, 472–481. [Google Scholar] [CrossRef] [PubMed]
  93. Liang, Z.-X.; Sun, J.-P.; Wang, P.; Tian, Q.; Yang, Z.; Chen, L.-A. Bone marrow-derived mesenchymal stem cells protect rats from endotoxin-induced acute lung injury. Chin. Med. J. 2011, 124, 2715–2722. [Google Scholar]
  94. Bonfield, T.L.; Koloze, M.; Lennon, D.P.; Zuchowski, B.; Yang, S.E.; Caplan, A.I. Human mesenchymal stem cells suppress chronic airway inflammation in the murine ovalbumin asthma model. Am. J. Physiol. Lung Cell Mol. Physiol. 2010, 299, L760–L770. [Google Scholar] [CrossRef] [PubMed]
  95. Nemeth, K.; Keane-Myers, A.; Brown, J.M.; Metcalfe, D.D.; Gorham, J.D.; Gorham, J.D.; Bundoc, V.G.; Bundoc, V.G.; Hodges, M.G.; Jelinek, I.; et al. Bone marrow stromal cells use TGF-beta to suppress allergic responses in a mouse model of ragweed-induced asthma. Proc. Natl. Acad. Sci. USA 2010, 107, 5652–5657. [Google Scholar] [CrossRef] [Green Version]
  96. Goodwin, M.; Sueblinvong, V.; Eisenhauer, P.; Ziats, N.P.; LeClair, L.; Poynter, M.E.; Steele, C.; Rincon, M.; Weiss, D.J. Bone marrow-derived mesenchymal stromal cells inhibit Th2-mediated allergic airways inflammation in mice. Stem Cells 2011, 29, 1137–1148. [Google Scholar] [CrossRef] [Green Version]
  97. van Haaften, T.; Byrne, R.; Bonnet, S.; Rochefort, G.Y.; Akabutu, J.; Bouchentouf, M.; Rey-Parra, G.J.; Galipeau, J.; Haromy, A.; Eaton, F.; et al. Airway delivery of mesenchymal stem cells prevents arrested alveolar growth in neonatal lung injury in rats. Am. J. Respir. Crit. Care Med. 2009, 180, 1131–1142. [Google Scholar] [CrossRef] [Green Version]
  98. Tropea, K.A.; Leder, E.; Aslam, M.; Lau, A.N.; Raiser, D.M.; Lee, J.-H.; Balasubramaniam, V.; Fredenburgh, L.E.; Alex Mitsialis, S.; Kourembanas, S.; et al. Bronchioalveolar stem cells increase after mesenchymal stromal cell treatment in a mouse model of bronchopulmonary dysplasia. Am. J. Physiol. Lung Cell Mol. Physiol. 2012, 302, L829–L837. [Google Scholar] [CrossRef]
  99. Ishizawa, K.; Kubo, H.; Yamada, M.; Kobayashi, S.; Numasaki, M.; Ueda, S.; Suzuki, T.; Sasaki, H. Bone marrow-derived cells contribute to lung regeneration after elastase-induced pulmonary emphysema. FEBS Lett. 2004, 556, 249–252. [Google Scholar] [CrossRef]
  100. Zhen, G.; Liu, H.; Gu, N.; Zhang, H.; Xu, Y.; Zhang, Z. Mesenchymal stem cells transplantation protects against rat pulmonary emphysema. Front. Biosci. 2008, 13, 3415–3422. [Google Scholar] [CrossRef] [Green Version]
  101. Aguilar, S.; Scotton, C.J.; McNulty, K.; Nye, E.; Stamp, G.; Laurent, G.; Bonnet, D.; Janes, S.M. Bone marrow stem cells expressing keratinocyte growth factor via an inducible lentivirus protects against bleomycin-induced pulmonary fibrosis. PLoS ONE 2009, 4, e8013. [Google Scholar] [CrossRef]
  102. Cargnoni, A.; Ressel, L.; Rossi, D.; Poli, A.; Arienti, D.; Lombardi, G.; Parolini, O. Conditioned medium from amniotic mesenchymal tissue cells reduces progression of bleomycin-induced lung fibrosis. Cytotherapy 2012, 14, 153–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Lee, S.-H.; Jang, A.-S.; Kim, Y.-E.; Cha, J.-Y.; Kim, T.-H.; Jung, S.; Park, S.-K.; Lee, Y.-K.; Won, J.-H.; Kim, Y.-H.; et al. Modulation of cytokine and nitric oxide by mesenchymal stem cell transfer in lung injury/fibrosis. Respir. Res. 2010, 11, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Moodley, Y.; Atienza, D.; Manuelpillai, U.; Samuel, C.S.; Tchongue, J.; Ilancheran, S.; Boyd, R.; Trounson, A. Human umbilical cord mesenchymal stem cells reduce fibrosis of bleomycin-induced lung injury. Am. J. Pathol. 2009, 175, 303–313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ortiz, L.A.; Dutreil, M.; Fattman, C.; Pandey, A.C.; Torres, G.; Go, K.; Phinney, D.G. Interleukin 1 receptor antagonist mediates the antiinflammatory and antifibrotic effect of mesenchymal stem cells during lung injury. Proc. Natl. Acad. Sci. USA 2007, 104, 11002–11007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Rojas, M.; Xu, J.; Woods, C.R.; Mora, A.L.; Spears, W.; Roman, J.; Brigham, K.L. Bone marrow-derived mesenchymal stem cells in repair of the injured lung. Am. J. Respir. Cell Mol. Biol. 2005, 33, 145–152. [Google Scholar] [CrossRef] [Green Version]
  107. Shigemura, N.; Okumura, M.; Mizuno, S.; Imanishi, Y.; Nakamura, T.; Sawa, Y. Autologous transplantation of adipose tissue-derived stromal cells ameliorates pulmonary emphysema. Am. J. Transplant. 2006, 6, 2592–2600. [Google Scholar] [CrossRef]
  108. Gonzalez-Rey, E.; Anderson, P.; González, M.A.; Rico, L.; Büscher, D.; Delgado, M. Human adult stem cells derived from adipose tissue protect against experimental colitis and sepsis. Gut 2009, 58, 929–939. [Google Scholar] [CrossRef]
  109. Yagi, H.; Soto-Gutierrez, A.; Navarro-Alvarez, N.; Nahmias, Y.; Goldwasser, Y.; Kitagawa, Y.; Tilles, A.W.; Tompkins, R.G.; Parekkadan, B.; Yarmush, M.L. Reactive bone marrow stromal cells attenuate systemic inflammation via sTNFR1. Mol. Ther. 2010, 18, 1857–1864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Curley, G.F.; Hayes, M.; Ansari, B.; Shaw, G.; Ryan, A.; Barry, F.; O’Brien, T.; O’Toole, D.; Laffey, J.G. Mesenchymal stem cells enhance recovery and repair following ventilator-induced lung injury in the rat. Thorax 2012, 67, 496–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Salem, H.K.; Thiemermann, C. Mesenchymal stromal cells: Current understanding and clinical status. Stem Cells 2010, 28, 585–596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Katsuda, T.; Kosaka, N.; Takeshita, F.; Ochiya, T. The therapeutic potential of mesenchymal stem cell-derived extracellular vesicles. Proteomics 2013, 13, 1637–1653. [Google Scholar] [CrossRef] [PubMed]
  113. Caretti, A.; Peli, V.; Colombo, M.; Zulueta, A. Lights and Shadows in the Use of Mesenchymal Stem Cells in Lung Inflammation, a Poorly Investigated Topic in Cystic Fibrosis. Cells 2019, 9, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Wong, A.P.; Dutly, A.E.; Sacher, A.; Lee, H.; Hwang, D.M.; Liu, M.; Keshavjee, S.; Hu, J.; Waddell, T.K. Targeted cell replacement with bone marrow cells for airway epithelial regeneration. Am. J. Physiol. Lung Cell Mol. Physiol. 2007, 293, L740–L752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Wong, A.P.; Keating, A.; Lu, W.-Y.; Duchesneau, P.; Wang, X.; Sacher, A.; Hu, J.; Waddell, T.K. Identification of a bone marrow-derived epithelial-like population capable of repopulating injured mouse airway epithelium. J. Clin. Investig. 2009, 119, 336–348. [Google Scholar] [CrossRef] [Green Version]
  116. Gilpin, S.E.; Lung, K.C.; Sato, M.; Singer, L.G.; Keshavjee, S.; Waddell, T.K. Altered progenitor cell and cytokine profiles in bronchiolitis obliterans syndrome. J. Heart Lung Transplant. 2012, 31, 222–228. [Google Scholar] [CrossRef]
  117. Duchesneau, P.; Wong, A.P.; Waddell, T.K. Optimization of targeted cell replacement therapy: A new approach for lung disease. Mol. Ther. 2010, 18, 1830–1836. [Google Scholar] [CrossRef]
  118. Wang, G.; Bunnell, B.A.; Painter, R.G.; Quiniones, B.C.; Tom, S.; Lanson, N.A.; Spees, J.L.; Bertucci, D.; Peister, A.; Weiss, D.J.; et al. Adult stem cells from bone marrow stroma differentiate into airway epithelial cells: Potential therapy for cystic fibrosis. Proc. Natl. Acad. Sci. USA 2005, 102, 186–191. [Google Scholar] [CrossRef] [Green Version]
  119. Paracchini, V.; Carbone, A.; Colombo, F.; Castellani, S.; Mazzucchelli, S.; Gioia, S.D.; Degiorgio, D.; Seia, M.; Porretti, L.; Colombo, C.; et al. Amniotic mesenchymal stem cells: A new source for hepatocyte-like cells and induction of CFTR expression by coculture with cystic fibrosis airway epithelial cells. J. Biomed. Biotechnol. 2012, 2012, 575471. [Google Scholar] [CrossRef]
  120. Carbone, A.; Castellani, S.; Favia, M.; Diana, A.; Paracchini, V.; Di Gioia, S.; Seia, M.; Casavola, V.; Colombo, C.; Conese, M. Correction of defective CFTR/ENaC function and tightness of cystic fibrosis airway epithelium by amniotic mesenchymal stromal (stem) cells. J. Cell. Mol. Med. 2014, 18, 1631–1643. [Google Scholar] [CrossRef]
  121. Sutton, M.T.; Fletcher, D.; Ghosh, S.K.; Weinberg, A.; van Heeckeren, R.; Kaur, S.; Sadeghi, Z.; Hijaz, A.; Reese, J.; Lazarus, H.M.; et al. Antimicrobial Properties of Mesenchymal Stem Cells: Therapeutic Potential for Cystic Fibrosis Infection, and Treatment. Stem Cells Int 2016, 2016, 5303048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Sutton, M.T.; Fletcher, D.; Episalla, N.; Auster, L.; Kaur, S.; Gwin, M.C.; Folz, M.; Velasquez, D.; Roy, V.; van Heeckeren, R.; et al. Mesenchymal Stem Cell Soluble Mediators and Cystic Fibrosis. J. Stem Cell Res. Ther 2017, 7. [Google Scholar] [CrossRef] [Green Version]
  123. Li, M.; Pascual, G.; Glass, C.K. Peroxisome proliferator-activated receptor gamma-dependent repression of the inducible nitric oxide synthase gene. Mol. Cell. Biol. 2000, 20, 4699–4707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Lee, C.; Mitsialis, S.A.; Aslam, M.; Vitali, S.H.; Vergadi, E.; Konstantinou, G.; Sdrimas, K.; Fernandez-Gonzalez, A.; Kourembanas, S. Exosomes mediate the cytoprotective action of mesenchymal stromal cells on hypoxia-induced pulmonary hypertension. Circulation 2012, 126, 2601–2611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Zhu, Y.-G.; Feng, X.-M.; Abbott, J.; Fang, X.-H.; Hao, Q.; Monsel, A.; Qu, J.-M.; Matthay, M.A.; Lee, J.W. Human mesenchymal stem cell microvesicles for treatment of Escherichia coli endotoxin-induced acute lung injury in mice. Stem Cells 2014, 32, 116–125. [Google Scholar] [CrossRef] [Green Version]
  126. Zhang, H.-C.; Liu, X.-B.; Huang, S.; Bi, X.-Y.; Wang, H.-X.; Xie, L.-X.; Wang, Y.-Q.; Cao, X.-F.; Lv, J.; Xiao, F.-J.; et al. Microvesicles derived from human umbilical cord mesenchymal stem cells stimulated by hypoxia promote angiogenesis both in vitro and in vivo. Stem Cells Dev. 2012, 21, 3289–3297. [Google Scholar] [CrossRef]
  127. Lai, R.C.; Arslan, F.; Lee, M.M.; Sze, N.S.K.; Choo, A.; Chen, T.S.; Salto-Tellez, M.; Timmers, L.; Lee, C.N.; El Oakley, R.M.; et al. Exosome secreted by MSC reduces myocardial ischemia/reperfusion injury. Stem Cell Res. 2010, 4, 214–222. [Google Scholar] [CrossRef] [Green Version]
  128. Zulueta, A.; Colombo, M.; Peli, V.; Falleni, M.; Tosi, D.; Ricciardi, M.; Baisi, A.; Bulfamante, G.; Chiaramonte, R.; Caretti, A. Lung mesenchymal stem cells-derived extracellular vesicles attenuate the inflammatory profile of Cystic Fibrosis epithelial cells. Cell. Signal 2018, 51, 110–118. [Google Scholar] [CrossRef]
  129. Guo, L.; Karoubi, G.; Duchesneau, P.; Shutova, M.V.; Sung, H.-K.; Tonge, P.; Bear, C.; Rogers, I.; Nagy, A.; Waddell, T.K. Generation of Induced Progenitor-like Cells from Mature Epithelial Cells Using Interrupted Reprogramming. Stem Cell Rep. 2017, 9, 1780–1795. [Google Scholar] [CrossRef] [Green Version]
  130. Guo, L.; Karoubi, G.; Duchesneau, P.; Aoki, F.G.; Shutova, M.V.; Rogers, I.; Nagy, A.; Waddell, T.K. Interrupted reprogramming of alveolar type II cells induces progenitor-like cells that ameliorate pulmonary fibrosis. NPJ Regen Med. 2018, 3, 14. [Google Scholar] [CrossRef]
  131. 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]
  132. Soh, B.S.; Zheng, D.; Li Yeo, J.S.; Yang, H.H.; Ng, S.Y.; Wong, L.H.; Zhang, W.; Li, P.; Nichane, M.; Asmat, A.; et al. CD166(pos) subpopulation from differentiated human ES and iPS cells support repair of acute lung injury. Mol. Ther. 2012, 20, 2335–2346. [Google Scholar] [CrossRef] [Green Version]
  133. Wang, R.; McCauley, K.B.; Kotton, D.N.; Hawkins, F. Differentiation of human airway-organoids from induced pluripotent stem cells (iPSCs). Methods Cell Biol. 2020, 159, 95–114. [Google Scholar] [PubMed]
  134. Shafa, M.; Ionescu, L.I.; Vadivel, A.; Collins, J.J.P.; Xu, L.; Zhong, S.; Kang, M.; de Caen, G.; Daneshmand, M.; Shi , J.; et al. Human induced pluripotent stem cell-derived lung progenitor and alveolar epithelial cells attenuate hyperoxia-induced lung injury. Cytotherapy 2018, 20, 108–125. [Google Scholar] [CrossRef] [PubMed]
  135. Su, V.Y.-F.; Yang, K.-Y.; Chiou, S.-H.; Chen, N.-J.; Mo, M.-H.; Lin, C.-S.; Wang, C.-T. Induced Pluripotent Stem Cells Regulate Triggering Receptor Expressed on Myeloid Cell-1 Expression and the p38 Mitogen-Activated Protein Kinase Pathway in Endotoxin-Induced Acute Lung Injury. Stem Cells 2019, 37, 631–639. [Google Scholar] [CrossRef] [PubMed]
  136. Yao, Y.; Fan, X.-L.; Jiang, D.; Zhang, Y.; Li, X.; Xu, Z.-B.; Fang, S.-B.; Chiu, S.; Tse, H.-F.; Lian, Q.; et al. Connexin 43-Mediated Mitochondrial Transfer of iPSC-MSCs Alleviates Asthma Inflammation. Stem Cell Rep. 2018, 11, 1120–1135. [Google Scholar] [CrossRef] [Green Version]
  137. Tamò, L.; Simillion, C.; Hibaoui, Y.; Feki, A.; Gugger, M.; Prasse, A.; Jäger, B.; Goldmann, T.; Geiser, T.; Gazdhar, A. Gene Network Analysis of Interstitial Macrophages After Treatment with Induced Pluripotent Stem Cells Secretome (iPSC-cm) in the Bleomycin Injured Rat Lung. Stem Cell Rev. Rep. 2018, 14, 412–424. [Google Scholar] [CrossRef] [Green Version]
  138. Mitchell, A.; Wanczyk, H.; Jensen, T.; Finck, C. Human induced pluripotent stem cells ameliorate hyperoxia-induced lung injury in a mouse model. Am. J. Transl. Res. 2020, 12, 292–307. [Google Scholar]
  139. de Miguel-Beriain, I. The ethics of stem cells revisited. Adv. Drug Deliv. Rev. 2015, 82-83, 176–180. [Google Scholar] [CrossRef]
  140. Ilic, D.; Ogilvie, C. Concise Review: Human Embryonic Stem Cells-What Have We Done? What Are We Doing? Where Are We Going? Stem Cells 2017, 35, 17–25. [Google Scholar] [CrossRef] [Green Version]
  141. Moodley, Y.; Thompson, P.; Warburton, D. Stem cells: A recapitulation of development. Respirology 2013, 18, 1167–1176. [Google Scholar] [CrossRef] [PubMed]
  142. Serra, M.; Alysandratos, K.-D.; Hawkins, F.; McCauley, K.B.; Jacob, A.; Choi, J.; Caballero, I.S.; Vedaie, M.; Kurmann, A.A.; Ikonomou, L.; et al. Pluripotent stem cell differentiation reveals distinct developmental pathways regulating lung- versus thyroid-lineage specification. Development 2017, 144, 3879–3893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Kimura, Y.; Shofuda, T.; Higuchi, Y.; Nagamori, I.; Oda, M.; Nakamori, M.; Onodera, M.; Kanematsu, D.; Yamamoto, A.; Katsuma, A.; et al. Human Genomic Safe Harbors and the Suicide Gene-Based Safeguard System for iPSC-Based Cell Therapy. Stem Cells Transl. Med. 2019, 8, 627–638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Ben-David, U.; Benvenisty, N. Chemical ablation of tumor-initiating human pluripotent stem cells. Nat. Protoc. 2014, 9, 729–740. [Google Scholar] [CrossRef]
  145. Tan, H.L.; Fong, W.J.; Lee, E.H.; Yap, M.; Choo, A. mAb 84, a cytotoxic antibody that kills undifferentiated human embryonic stem cells via oncosis. Stem Cells 2009, 27, 1792–1801. [Google Scholar] [CrossRef]
  146. Blum, B.; Bar-Nur, O.; Golan-Lev, T.; Benvenisty, N. The anti-apoptotic gene survivin contributes to teratoma formation by human embryonic stem cells. Nat. Biotechnol. 2009, 27, 281–287. [Google Scholar] [CrossRef]
  147. Rong, Z.; Fu, X.; Wang, M.; Xu, Y. A scalable approach to prevent teratoma formation of human embryonic stem cells. J. Biol. Chem. 2012, 287, 32338–32345. [Google Scholar] [CrossRef] [Green Version]
  148. Lee, M.-O.; Moon, S.H.; Jeong, H.-C.; Yi, J.-Y.; Lee, T.-H.; Shim, S.H.; Rhee, Y.-H.; Lee, S.-H.; Oh, S.-J.; Lee, M.-Y.; et al. Inhibition of pluripotent stem cell-derived teratoma formation by small molecules. Proc. Natl. Acad. Sci. USA 2013, 110, E3281–E3290. [Google Scholar] [CrossRef] [Green Version]
  149. Liang, Q.; Monetti, C.; Shutova, M.V.; Neely, E.J.; Hacibekiroglu, S.; Yang, H.; Kim, C.; Zhang, P.; Li, C.; Nagy, K.; et al. Linking a cell-division gene and a suicide gene to define and improve cell therapy safety. Nature 2018, 563, 701–704. [Google Scholar] [CrossRef]
  150. Lanza, R.; Russell, D.W.; Nagy, A. Engineering universal cells that evade immune detection. Nat. Rev. Immunol. 2019, 19, 723–733. [Google Scholar] [CrossRef]
  151. Taylor, C.J.; Peacock, S.; Chaudhry, A.N.; Bradley, J.A.; Bolton, E.M. Generating an iPSC bank for HLA-matched tissue transplantation based on known donor and recipient HLA types. Cell Stem Cell 2012, 11, 147–152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Deuse, T.; Hu, X.; Gravina, A.; Wang, D.; Tediashvili, G.; De, C.; Thayer, W.O.; Wahl, A.; Garcia, J.V.; Reichenspurner, H.; et al. Hypoimmunogenic derivatives of induced pluripotent stem cells evade immune rejection in fully immunocompetent allogeneic recipients. Nat. Biotechnol. 2019, 37, 252–258. [Google Scholar] [CrossRef] [PubMed]
  153. Riolobos, L.; Hirata, R.K.; Turtle, C.J.; Wang, P.-R.; Gornalusse, G.G.; Zavajlevski, M.; Riddell, S.R.; Russell, D.W. HLA engineering of human pluripotent stem cells. Mol. Ther. 2013, 21, 1232–1241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Rong, Z.; Wang, M.; Hu, Z.; Stradner, M.; Zhu, S.; Kong, H.; Yi, H.; Goldrath, A.; Yang, Y.-G.; Xu, Y.; et al. An effective approach to prevent immune rejection of human ESC-derived allografts. Cell Stem Cell 2014, 14, 121–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Vaidyanathan, S.; Salahudeen, A.A.; Sellers, Z.M.; Bravo, D.T.; Choi, S.S.; Batish, A.; Le, W.; Baik, R.; Kaushik, M.P.; Galper, N.; et al. High-Efficiency, Selection-free Gene Repair in Airway Stem Cells from Cystic Fibrosis Patients Rescues CFTR Function in Differentiated Epithelia. Cell Stem Cell 2020, 26, 161–171.e4. [Google Scholar] [CrossRef] [PubMed]
  156. Hsu, P.D.; Scott, D.A.; Weinstein, J.A.; Ran, F.A.; Konermann, S.; Agarwala, V.; Li, Y.; Fine, E.J.; Wu, X.; Shalem, O.; et al. DNA targeting specificity of RNA-guided Cas9 nucleases. Nat. Biotechnol. 2013, 31, 827–832. [Google Scholar] [CrossRef]
  157. Lin, Y.; Cradick, T.J.; Brown, M.T.; Deshmukh, H.; Ranjan, P.; Sarode, N.; Wile, B.M.; Vertino, P.M.; Stewart, F.J.; Bao, G. CRISPR/Cas9 systems have off-target activity with insertions or deletions between target DNA and guide RNA sequences. Nucleic Acids Res. 2014, 42, 7473–7485. [Google Scholar] [CrossRef] [Green Version]
  158. Qadir, M.M.F.; Álvarez-Cubela, S.; Belle, K.; Sapir, T.; Messaggio, F.; Johnson, K.B.; Umland, O.; Hardin, D.; Klein, D.; Pérez-Álvarez, I.; et al. A Double Fail-Safe Approach to Prevent Tumorigenesis and Select Pancreatic β Cells from Human Embryonic Stem Cells. Stem Cell Rep. 2019, 12, 611–623. [Google Scholar] [CrossRef] [Green Version]
  159. Rosen, C.; Shezen, E.; Aronovich, A.; Klionsky, Y.Z.; Yaakov, Y.; Assayag, M.; Biton, I.E.; Tal, O.; Shakhar, G.; Ben-Hur, H.; et al. Preconditioning allows engraftment of mouse and human embryonic lung cells, enabling lung repair in mice. Nat. Med. 2015, 21, 869–879. [Google Scholar] [CrossRef]
  160. Farrow, N.; Cmielewski, P.; Donnelley, M.; Rout-Pitt, N.; Moodley, Y.; Bertoncello, I.; Parsons, D. Epithelial disruption: A new paradigm enabling human airway stem cell transplantation. Stem Cell Res. Ther. 2018, 9, 153–158. [Google Scholar] [CrossRef] [Green Version]
  161. Shiraishi, K.; Shichino, S.; Tsukui, T.; Hashimoto, S.; Ueha, S.; Matsushima, K. Engraftment and proliferation potential of embryonic lung tissue cells in irradiated mice with emphysema. Sci. Rep. 2019, 9, 3657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Cypel, M.; Yeung, J.C.; Liu, M.; Anraku, M.; Chen, F.; Karolak, W.; Sato, M.; Laratta, J.; Azad, S.; Madonik, M.; et al. Normothermic ex vivo lung perfusion in clinical lung transplantation. N. Engl. J. Med. 2011, 364, 1431–1440. [Google Scholar] [CrossRef] [Green Version]
  163. Nichane, M.; Javed, A.; Sivakamasundari, V.; Ganesan, M.; Ang, L.T.; Kraus, P.; Lufkin, T.; Loh, K.M.; Lim, B. Isolation and 3D expansion of multipotent Sox9+ mouse lung progenitors. Nat. Methods 2017, 14, 1205–1212. [Google Scholar] [CrossRef] [PubMed]
  164. Vaughan, A.E.; Brumwell, A.N.; Xi, Y.; Gotts, J.E.; Brownfield, D.G.; Treutlein, B.; Tan, K.; Tan, V.; Liu, F.C.; Looney, M.R.; et al. Lineage-negative progenitors mobilize to regenerate lung epithelium after major injury. Nature 2015, 517, 621–625. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic of cell-based replacement therapy for Cystic Fibrosis. Cell-replacement therapy in which defective airway epithelium (1) is ablated via injury to the airways thereby creating a niche for engraftment (2). Corrected cells are then delivered into the airways, localize and engraft in the exposed niche (3) and restore functional epithelium (4). Figure created with BioRender.com.
Figure 1. Schematic of cell-based replacement therapy for Cystic Fibrosis. Cell-replacement therapy in which defective airway epithelium (1) is ablated via injury to the airways thereby creating a niche for engraftment (2). Corrected cells are then delivered into the airways, localize and engraft in the exposed niche (3) and restore functional epithelium (4). Figure created with BioRender.com.
Ijms 21 05219 g001
Figure 2. Schematic representation of the Induced progenitor-like cells (iPL) process and their utility in cell therapy for Cystic Fibrosis (CF). iPL are generated from mature club cells isolated from a CF patient via transient reprogramming with Oct4, Sox2, Klf4, and c-Myc (OKSM). Gene corrected iPL cells are then transplanted back to the preconditioned recipient airways where they will engraft and integrate into the airways and restore the epithelium. Figure created with BioRender.com.
Figure 2. Schematic representation of the Induced progenitor-like cells (iPL) process and their utility in cell therapy for Cystic Fibrosis (CF). iPL are generated from mature club cells isolated from a CF patient via transient reprogramming with Oct4, Sox2, Klf4, and c-Myc (OKSM). Gene corrected iPL cells are then transplanted back to the preconditioned recipient airways where they will engraft and integrate into the airways and restore the epithelium. Figure created with BioRender.com.
Ijms 21 05219 g002
Figure 3. Derivation of proximal airway epithelial cells from induced pluripotent stem (iPS) cells. Schematic representation of the embryologically guided stepwise approach to generate mature proximal epithelial cells. The first requirement is the generation of definitive endoderm (equivalent to cells at approximately 3 weeks post conception) from iPS cells (equivalent to the cells found in the inner cell mass of the blastocyst at day 6). The second stage is differentiation to anterior foregut endoderm (equivalent to cells at approximately 4 weeks post conception). Cells are then differentiated via ventralization to NKX2.1+ lung progenitors (equivalent to cells at approximately 5–6 weeks post conception). Proximal airway lineage and maturation is then induced using 3D culture conditions and air–liquid-interface (ALI) culture. Figure created with BioRender.com.
Figure 3. Derivation of proximal airway epithelial cells from induced pluripotent stem (iPS) cells. Schematic representation of the embryologically guided stepwise approach to generate mature proximal epithelial cells. The first requirement is the generation of definitive endoderm (equivalent to cells at approximately 3 weeks post conception) from iPS cells (equivalent to the cells found in the inner cell mass of the blastocyst at day 6). The second stage is differentiation to anterior foregut endoderm (equivalent to cells at approximately 4 weeks post conception). Cells are then differentiated via ventralization to NKX2.1+ lung progenitors (equivalent to cells at approximately 5–6 weeks post conception). Proximal airway lineage and maturation is then induced using 3D culture conditions and air–liquid-interface (ALI) culture. Figure created with BioRender.com.
Ijms 21 05219 g003
Figure 4. Schematic representation of cell-based therapeutic approaches using induced pluripotent stem cells. In autologous iPS cell therapy, somatic cells are isolated from the CF patient, expanded and reprogrammed to produce patient-specific induced pluripotent cells. These cells are subsequently differentiated to proximal airway epithelium, corrected for the genetic defect to achieve normal airway epithelium and transplanted back to the patient. In allogeneic iPS cell therapy, a universal iPS line (from a healthy source) which has been genetically modified for safety and immunogenicity is used to generate proximal airway epithelium which is transplanted to CF patients. Figure created with BioRender.com.
Figure 4. Schematic representation of cell-based therapeutic approaches using induced pluripotent stem cells. In autologous iPS cell therapy, somatic cells are isolated from the CF patient, expanded and reprogrammed to produce patient-specific induced pluripotent cells. These cells are subsequently differentiated to proximal airway epithelium, corrected for the genetic defect to achieve normal airway epithelium and transplanted back to the patient. In allogeneic iPS cell therapy, a universal iPS line (from a healthy source) which has been genetically modified for safety and immunogenicity is used to generate proximal airway epithelium which is transplanted to CF patients. Figure created with BioRender.com.
Ijms 21 05219 g004

Share and Cite

MDPI and ACS Style

Duchesneau, P.; Waddell, T.K.; Karoubi, G. Cell-Based Therapeutic Approaches for Cystic Fibrosis. Int. J. Mol. Sci. 2020, 21, 5219. https://doi.org/10.3390/ijms21155219

AMA Style

Duchesneau P, Waddell TK, Karoubi G. Cell-Based Therapeutic Approaches for Cystic Fibrosis. International Journal of Molecular Sciences. 2020; 21(15):5219. https://doi.org/10.3390/ijms21155219

Chicago/Turabian Style

Duchesneau, Pascal, Thomas K. Waddell, and Golnaz Karoubi. 2020. "Cell-Based Therapeutic Approaches for Cystic Fibrosis" International Journal of Molecular Sciences 21, no. 15: 5219. https://doi.org/10.3390/ijms21155219

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