Abstract
Delivery of genetic material to tissues in vivo is an important technique used in research settings and is the foundation upon which clinical gene therapy is built. The lung is a prime target for gene delivery due to a host of genetic, acquired, and infectious diseases that manifest themselves there, resulting in many pathologies. However, the in vivo delivery of genetic material to the lung remains a practical problem clinically and is considered the major obstacle needed to be overcome for gene therapy. Currently there are four main strategies for in vivo gene delivery to the lung: viral vectors, liposomes, nanoparticles, and electroporation. Viral delivery uses several different genetically modified viruses that enter the cell and express desired genes that have been inserted to the viral genome. Liposomes use combinations of charged and neutral lipids that can encapsulate genetic cargo and enter cells through endogenous mechanisms, thereby delivering their cargoes. Nanoparticles are defined by their size (typically less than 100 nm) and are made up of many different classes of building blocks, including biological and synthetic polymers, cell penetrant and other peptides, and dendrimers, that also enter cells through endogenous mechanisms. Electroporation uses mild to moderate electrical pulses to create pores in the cell membrane through which delivered genetic material can enter a cell. An emerging fifth category, exosomes and extracellular vesicles, may have advantages of both viral and non-viral approaches. These extracellular vesicles bud from cellular membranes containing receptors and ligands that may aid cell targeting and which can be loaded with genetic material for efficient transfer. Each of these vectors can be used for different gene delivery applications based on mechanisms of action, side-effects, and other factors, and their use in the lung and possible clinical considerations is the primary focus of this review.
Keywords: Lung, gene delivery, viral vectors, liposomes, nanoparticles, electroporation, exosomes
Impact statement
This work provides an update on basic, pre-clinical, and clinical gene delivery to the lung, highlighting improvements and remaining challenges to effective in vivo gene therapy. This provides important information on questions and problems in the field that should be addressed to further translation of work from the laboratory to the clinic. We further highlight new/undeveloped techniques of interest to the field. The information contained within has not been presented from this viewpoint or format previously, to our knowledge.
The current state of gene therapy
Clinical gene delivery is a recent development, with only 17 FDA-approved therapies available as of August 2019. Most of these FDA-approved therapies involve the reprogramming of autologous cells externally before transplant of the reprogrammed cells into a patient. The majority of these approaches are for immunotherapy for various cancers, and many employ CAR-T cells. Indeed, only two current FDA-approved drugs are for gene therapy directly, Luxturna and Zolgensma, using viruses to transfer genes to treat congenital blindness and spinal muscular atrophy, respectively. However, none of the approved gene therapies currently target the lung, although over 180 Phase I, II, and III clinical trials have been completed and another 250 are currently progressing (ClinicalTrials.gov, 2019). Transplantation strategies come with many potential problems due to pre-existing immunity and failed engraftment. Some of these therapies may also face ethical challenges due to questions concerning the sources of the transplanted cells used.1–4 Additionally, prices for patients are upwards of hundreds of thousands to millions of dollars for these early cell-based gene therapies.5–8 Given these drawbacks, several different methods are being developed in order to deliver genetic therapies in vivo without having to transplant cells. These methods aim to decrease cost for the patient and reduce the potential of adverse effects.
The main challenges of in vivo gene delivery include off-target effects of the vector itself or the genetic material and delivery efficiency. Early approaches for gene therapy were largely focused on gene overexpression or methods to repair a mutant gene using viral or non-viral delivery strategies. Inefficiency of gene transfer, immunological responses, and non-specificity of cell targeting are just a few of the problems associated with viral approaches for gene delivery. By contrast, many methods of non-viral delivery have been less robust, yielding lower levels of transfection in vivo. When successful, it has even been shown that overexpressing plasmids sometimes can outcompete host gene expression in a competitive manner, making it difficult to modulate appropriate levels of transgene expression. Early gene correction technology focused on zinc-finger nucleases and TALENS, though these are now being replaced by more efficient CRISPR-Cas technologies. As gene therapy has evolved, it has been realized that more than simply overexpression or gene replacement can be done. One of these approaches have been to knockdown aberrant or other gene expression utilizing RNA interference (RNAi) technologies. Many different nucleic acids can be used for RNAi, including siRNA, shRNA, bifunctional shRNA, lncRNA, and miRNA. The different nucleic acid species used can have large differences in off-target effects and the duration and degree of target gene knockdown; both siRNA and shRNA are highly specific for a given target gene, while lncRNA and miRNA are more promiscuous due to less stringent homology needed for binding to their targets and thus can target multiple genes in entire pathways. While many RNAi approaches have used direct transfer of modified RNAs, both shRNA and miRNA can be expressed in target cells via transferred plasmids for longer term expression as well.9–11 Functional delivery of all of these genetic materials in vivo has remained problematic due to issues with vector off-target effects, delivery efficiency, and perceived and real safety issues.
Disease targets in the lung
The lung is a prime target for gene delivery due to its importance in maintaining homeostasis. Abnormal lung function is a common cause of death and is strongly correlated with the onset and severity of other pathologies. Indeed, the moto of the American Lung Association is “when you can’t breathe, nothing else matters”. The lung is the primary organ affected in many acute and chronic pathologies including cystic fibrosis, acute respiratory distress syndrome (ARDS), familial emphysema, pulmonary fibrosis, cancer, and numerous bacterial and viral infections. For example, cystic fibrosis (CF) is caused by a familial defect in CFTR, a gene responsible for chloride ion transport and necessary for normal function of the lung, gut, and pancreas. Abnormal secretion of mucus in the lung as a result of altered CFTR localization and activity decreases mucociliary clearance in the lung, decreases oxygenation, allows persistent pathogenic bacterial colonization, and eventually leads to death.12,13 Several recent drugs (e.g. Trikafta from Vertex Pharmaceuticals) have been developed that tremendously improve treatment for some genetic variants of CFTR by improving protein folding and channel activity, but many mutations are not effectively treated in this manner, especially nonsense mutations of the gene in which no full length CFTR is produced. Treatment for such mutations, which account for approximately 10% of all cystic fibrosis patients, thus necessitates some form of genetic therapy.12, 13,16 Another example is ARDS, which has a high mortality rate and no therapeutic approach besides basic symptomatic and supportive care strategies. ARDS is caused by many different insults, most commonly sepsis, lung trauma, pneumonia, acid or toxic gas inhalation, and viral infections, most recently highlighted by the COVID-19 pandemic, all of which result in the breakdown of the alveolar-capillary barrier, increased systemic inflammation, reduced gas exchange, hypoxemia, and ultimately multi-organ failure and is usually concurrent with other trauma or pathology making clinical care and treatment more difficult.14,15 Again, as with cystic fibrosis, since many traditional drugs have failed to show any activity against ARDS, a genetic approach seems plausible. Familial emphysema results from a deficiency of alpha-1 antitrypsin, causing widespread damage to and simplification of alveoli (causing greatly reduced gas exchange and lung function) requiring plasma or serum perfusions weekly-monthly.17--19 Like cystic fibrosis, this single gene mutation (or deletion) has long been a prime target for gene therapy approaches. Chronic obstructive pulmonary disease (COPD) is a largely smoking-associated disorder that also has aberrant alpha-1 antitrypsin activity, simplified alveolar structure, greatly reduced lung function, and similarly has no treatment currently.20--22 Interstitial lung disease, more commonly known as pulmonary fibrosis, is a broad category of disease where lung tissue undergoes a fibrotic transformation in response to injury or disease caused by perturbations of cellular and genetic signaling. As for ARDS and COPD, no single gene is responsible for the disease and as such, treatment will require a complex genetic approach in order to generate a therapeutic response.14,15,21,22 The lung is also a common site of cancer metastasis and has a relatively high possibility of primary tumors as well.23,24 Finally, due to the lung’s interaction with the environment, viral, bacterial, and fungal infections are also common and often associated with devastating consequences. Significant examples include tuberculosis, invasive pulmonary aspergillosis, and bacterial and viral pneumonia. These latter two can and often do lead to ARDS, as the world is currently experiencing with the novel COVID-19 coronavirus pandemic. In all cases and for all of these diseases, genetic therapies are of great importance due to lack of effective and widely available pharmacological therapeutics and the limitations of clinical care.
Barriers to gene delivery in the lung
While the lung is well suited to many forms of gene and drug delivery primarily due to its accessibility (via nose or mouth) and extremely large epithelial surface area, a number of physical and biological barriers are present (Figure 1). Any lung-targeted therapy must first pass a number of innate barriers, including mucus, pulmonary surfactant, ciliary beating and clearance, airway branching, innate immune responses, and local inflammation. Even in healthy individuals, the mucus and surfactant present in the lung has small nanosized pores which can impair or even prevent large molecules from passing through the lining fluid layer to the target cells below. This is even more of an issue in individuals with certain pathologies such as cystic fibrosis or asthma where mucus hypersecretion is a hallmark of the diseases. In the healthy lung where lining fluid is not abnormally viscous, the cilia constantly beat to clear the airways of unwanted particles, bacteria, and viruses. While this is beneficial for protecting the lung, the clearance can also cause reduced gene delivery by the same mechanisms.25,26 Airway branching also makes it difficult for drugs that cannot be aerosolized to target more than limited portions of the lung. Although bolus delivery of drugs and nanoparticles has been shown to mediate relatively even distribution to the deep lung in animal models,27 this is dependent on bolus volume and rate of administration, both of which are often not amenable to humans. Further, phagocytosis of delivered particles and by resident lung macrophages and their activation to induce widespread pulmonary and systemic inflammation can make drug design more challenging. This coupled with robust inflammatory responses caused by the induction and release of pro-inflammatory cytokines and chemokines by the pulmonary epithelium and endothelium can cause significant, and even deadly, inflammation, greatly limiting any effects of gene transfer. Lastly, if cells other than the epithelium are to be targets for gene delivery following airway delivery, the epithelium itself becomes a major barrier. Indeed, most gene delivery agents transfect or transduce only those cells in which they come in contact with, leaving the sub-epithelial cells largely untouched by the delivery agent unless damage to the epithelial lining occurs. Thus, when most viruses or non-viral vectors are used, no gene transfer to the endothelium, fibroblasts, smooth muscle, or other subepithelial cell is obtained. Bypassing these various barriers and targeting specific cells is therefore a large goal of gene delivery technology.
Figure 1.
Barriers to gene delivery in the lung. A number of physical, chemical, and physiological barriers for gene and drug delivery in the lung are shown.
Viral vectors
Viral vectors are one of the oldest non-chemical methods of gene delivery. First developed in the 1970s, viral vectors have largely remained the forerunners in gene therapy development. Initially, the viral vectors used for gene delivery were integrating retroviruses. While no longer commonly used for direct in vivo delivery, they are used extensively for ex vivo transduction of T-cells for cell therapies. Currently, adenovirus, adeno-associated virus (AAV), and lentiviruses are the most common vectors for viral gene delivery. In vivo use of viral vectors has had difficulty with immune responses, and even FDA-approved viral gene therapies require package insert warnings of possible cytokine release storm, a potentially lethal overactive immune response, as a side effect as well as other types of possible damage. Newer designs of viral gene delivery vectors have tried to limit their immunogenicity and replication capabilities to reduce these possible side effects.
The greatest advantage of all viral vectors is their inherent infectivity, or their ability to enter target cells, deliver cargo contained within the viral capsid, and lead to highly efficient gene delivery and expression (Figure 2, Tables 1 and 2). Due to the physiochemical properties of many capsid proteins and the small size of viral particles, viruses can often overcome several of the physical barriers of the lung, including mucus, surfactant, ciliary clearance, and airway branching. Viruses, such as retroviruses and lentiviruses, have innate mechanisms to integrate into the genome, allowing for long-term gene expression of the integrated transgene(s). The benefit or detriment of integration is context-dependent on whether sustained long-term or transient short-term expression is desired. Integration drives long-term expression, while a lack of integration for viruses most often provides short-term expression. In the treatment of inherited diseases, like cystic fibrosis or familial emphysema, integration for long-term expression may be preferred since long-term gene expression/replacement is needed to treat the disease. In contrast, an acute pathology, such as ARDS, is better treated in a transient manner, since the disease itself is also transient. This avoids potential side effects from integration and long-term upregulation or downregulation of pathways. A variety of endogenous viral capabilities allow vector choice to be made based on the type of pathology being treated and its specific needs. Due to these characteristics, viral vectors are already in use for several FDA-approved therapies for both short and long-term gene expression. The three most common viral vectors currently used, adenovirus, AAV, and lentiviruses, are each briefly discussed below.
Figure 2.
Relationship between gene therapy vectors and delivery barriers in the lung. Properties and abilities to overcome various delivery barriers in the lung are shown for the major classes of viral and non-viral methods of gene delivery (liposomes, nanoparticles, electroporation).
Table 1.
Vectors for gene delivery and delivery barriers in the lung.
|
Viral |
M Non-viral |
||||||
|---|---|---|---|---|---|---|---|
| Delivery barriers | Adenovirus | AAV | Lentivirus | Liposomes | Dendrimer nanoparticles | Cell penetrating peptide nanoparticles | Electroporation |
| Size | 80–100 nm | 20 nm | 80–100 nm | ∼30–1000 nm | Up to 100 nm | ∼5 nm | Size of DNA or RNA cargo |
| Dispersion | Yes | Yes | Yes | Yes | Size- and delivery-dependent | Delivery-dependent | Delivery-dependent |
| Removed by mucociliary clearance | Low | Low | Low | Some | Low | Low | No |
| Pre-existing immunity | Yes; serotype-dependent | Yes; serotype-dependent | No | No | Formulation-dependent | No | |
| Penetration of mucus/surfactant | serotype-dependent | serotype-dependent | Yes | Low | Good | Good | Good |
| Transfection of sub-epithelial cells | No | No | No | No | ? | ? | Yes |
Table 2.
Common gene delivery vectors used in the lung.
|
Viral |
Nonviral |
||||||
|---|---|---|---|---|---|---|---|
| Adenovirus | AAV | Lentivirus | Liposomes | Dendrimer nanoparticles | Cell penetrating peptide nanoparticles | Electroporation | |
| Genome size | 26–48 Kb | 4.7 Kb | 9–10 Kb | N/A | |||
| Size of transgene | 8 Kb max | 4.5 Kb max | 15 Kb max | No max; size proportional to 1/efficiency | Formula-dependent | Unknown | No max |
| RNAi (siRNA, miRNA, shRNA) | Yesa | Yes | Yes | Yes | Yes | siRNA and mi-RNA | Yes |
| Integrationb | Very low | Very low | Yes | No | No | No | No |
| Long-term expression | Uncommon | Yes | Yes | Dependent on promoter used | |||
| Exogenous material | Yes | Yes | Yes | No | No | No | No |
| Multiple deliveries | If serotype changes | If serotype changes | Unknown | yes | |||
aAdenovirus endogenously interferes with RNAi machinery (M. Anderson 2005, Virology).
bWithout additional factors such as transposases, for laboratory strains.
Adenovirus
Adenovirus is a non-enveloped, dsDNA virus with a capsid of approximately 80–100nm in size. Modified adenoviral genomes used for gene transfer are ∼30kb and can deliver ∼8kb of recombinant DNA.28 Wild-type adenoviruses cause a transient infection of the pulmonary tract in immunocompetent hosts. The transiency of these infections is due to high immunogenicity of the virus itself, both against capsid proteins and virus-encoded regulatory and replication proteins. Modifications to the adenoviral genome, removing most or all replication capability and immunogenic portions of the capsid proteins, have become standard when using helper-dependent and late generation adenoviral vectors in vivo. However, even with these modifications, most adenovirus-driven expression still lasts only one to two weeks in vivo due to a lack of integration and eventual immune clearance.29–31 While adenovirus appears to be a vector of choice for pulmonary gene therapy in the laboratory and has gene transfer efficiencies of almost 95% in vitro, the values are usually much less in vivo and require upwards of 109 to 1010 plaque forming units delivered intratracheally in the rat or mouse. Consequently, cell damage and inflammation are frequently observed, even with late generation helper-dependent viruses. Additionally, the major receptor for adenovirus sits in the basolateral membrane of the airway epithelium, requiring barrier disruption in order for viral transduction to occur.21,22 This makes its widespread clinical use for most diseases in the lung doubtful. Additionally, pre-existing immunity to many adenoviral serotypes limits their use for gene transfer in many patients and remains one of the greatest obstacles to its widespread clinical use.29–31 However, despite these issues, numerous studies have been published using adenoviral vectors for gene delivery to the lungs of animals due to the high levels of transgene expression obtained and have allowed researchers to test a variety of different genes in various disease models, including ARDS, cystic fibrosis, alpha-1 anti-trypsin deficiency, non-small cell lung carcinoma, and surfactant protein deficiencies.32,33 Further, this immunogenicity and transient expression have been fortuitous for using adenoviruses for in vivo tumor ablation and for vaccine development, where immunogenicity is advantageous.34 Some adenoviral serotypes can even specifically target cancer cells due to the increased glycosylation or polysialic acid on the tumor cells and a preference for binding to them by the vector.35,36 This type of cancer targeting can be used for direct lysis or modification of the cancer cells to improve other means of therapeutic targeting. Priming of the immune response by the vector itself can be desired for effective vaccination as it helps generate a memory response, reducing or eliminating the number of booster shots needed. In terms of human use, phase I trials of adenoviral vectors expressing TNF-alpha or interferon beta that target cancer cells in the lung for mesothelioma have been performed and showed some benefit, reducing tumor burden and increasing overall survival length in epithelial types of mesothelioma.37 Adenovirus is also used in FDA-approved therapies, where it is used ex vivo to deliver gene modification systems (such as CRISPR and sgRNA) to autologous cells to modify them for transplant, further expanding usage potential.38,39 However, despite its many advantages for high level gene transfer and expression in vivo, its use for vaccination, and studies in cancer settings using suicide gene therapy in both lab animals and humans, it has met relatively limited clinical success.33,40
Adeno-associated virus
Adeno-associated virus (AAV)-based vectors are much less inflammatory, common alternatives to adenoviral vectors. AAVs have greatly reduced pathogenicity compared to adenovirus, causing only a mild inflammatory response. This coupled with their ability to provide long-term gene expression without integration has led to their position in the gene therapy arsenal. AAV has a capsid that is only 20 nm in size, but can package and deliver ∼4.5 kb of recombinant DNA, almost the same size as the full AAV genome.41,42 Unlike adenovirus, AAV has a linear, single-stranded DNA genome. First discovered in the 1960s, wild-type AAV primarily integrates at a specific locus in chromosome 19. However, recombinant lab strains have reduced integration to 0.1% and any integration occurs at random sites throughout the genome. AAV remains in a lysogenic cycle forming an episomal, plasmid-like structure in the cytoplasm if the cell is not infected with a helper virus (such as adenovirus). Depending on the helper virus or helper viral proteins used, AAV can remain lysogenic, integrate into the genome, or enter an infectious lytic cycle. Reduced, but not non-existent, immunity and concern for integration has made AAV a favorable research tool. However, these same characteristics make it more difficult to use in cases where integration would be preferred for stable long-term expression. AAV helper viruses are also prevalent in the environment, making their use clinically less ideal than they would be otherwise.30,43 These environmental helper viruses could lead to unintentional gene integration or activate a lysogenic cycle and immune clearance of delivered AAV. Further, various levels of antibody-mediated immunity to most AAV serotypes are present throughout the population, and coupled with antibody responses generated following administration of recombinant AAV vectors, have limited repeat administration of the vectors to achieve long-lasting gene expression.33
While AAV serotypes have been used extensively in laboratory and pre-clinical studies in the lungs of multiple animal models to effectively delivery transgenes to a number of organs including the liver, eye, the CNS, and skeletal muscle, these vectors have been less successful in the lung.33,44 Nonetheless, AAV has been used with favorable results in vivo to treat CF in a pig CF model.43,45,46 An unexpected advantage of at least some serotypes of AAV, notably AAV6, is that this virus appears to be able to penetrate mucus with increased diffusion rates and distribution both in cultured cells and in mice with airway mucus obstruction, a property that could be very advantageous for CF therapies.47 Outside of the lung, two gene therapies using AAV as a vector have been FDA-approved, Luxturna and Zolgensma, to treat retinal dystrophy and spinal muscular atrophy, respectively. Luxturna uses a recombinant AAV2 to correct a mutation in RPE65 and Zolgensma uses an AAV9 capsid to deliver a normal SMN1 gene to motor neurons. Both are one dose only and carry hefty price tags in addition to caveats of childhood treatment and low disease severity.48–52
Lentivirus
Lentiviruses are positive-sense strand RNA viruses that form an 80–100 nm enveloped virion and have a delivery capacity of at least 15 kb. These viruses, like other retroviruses, contain reverse transcriptase to convert the RNA genome to a double-stranded DNA intermediate and integrase for its integration into the host genome.53 Because of its ability to integrate into the host genome, it allows for long-term gene expression and passage to any daughter cells. Perhaps the major advantage of lentiviruses over other retroviruses is their ability to successfully infect and transduce non-dividing cells. This allows them to transduce quiescent cells, including several critical targets, including T-cells and terminally differentiated somatic cells. Taken together, these abilities have made lentiviruses among the most-favored viral vectors for gene therapy. However, a major drawback to lentivirus use is also tied to its integration: lentivirus is able to integrate into germline cells making undesired longitudinal transfer to progeny a possibility. Further, the association of the lentiviral genus with HIV (and other immunodeficiency viruses) is not psychologically ideal.30,54 Perceived association with immunodeficiency viruses may affect the ease of clinical deployment due to patient concerns. The association with HIV is also not completely without merit; many lentiviral vectors used in research are in fact derived from HIV after removal of replication and virulence factors. As a result, research using lentivirus still requires protocols and personal protective equipment at biosafety level 2 to prevent the risk of infection to personnel.55,56 Affecting germline cells makes possible off-target and safety issues a matter for future generations as well. Further, since lentiviruses preferentially integrate in active gene sites, they have higher potential to affect normal cellular functions including oncogenic concerns.30,54 The exact factors through which lentivirus integration sites are determined are being studied and altered to make this process conducive to inactive gene integration, which could reduce the concerns associated with active site integration.
Of the viral vectors, lentivirus and similar retrovirus seems to be closest to widespread in vivo clinical deployment. The innate integration capability of lentivirus and other retrovirus has been key in developing chimeric antigen receptor T cells used in FDA-approved cell anti-cancer therapies (Kymriah and Yescarta).50–52 Further, lentiviral vectors are in phase 1/2 trials to treat many gene defects via ex vivo stem cell transplant. Stem cells harvested have a gene correction or defect-correcting alteration made and then are transplanted into an affected patient after ablation of the patient’s resident cells. Specific examples of this transplant approach include Duchenne muscular dystrophy,53 leukodystrophies,57,58 and beta-thalassemia.59 Lentiviral vectors have also been used effectively for gene transfer directly in the lung, primarily in small animal models. Simian immunodeficiency virus (SIV)-based vectors have been used to overexpress factor VIII and alpha-1 antitrypsin in the mouse lung following intratracheal administration, essentially turning the lungs into a bioreactor for protein overexpression.60 Other lentiviral vectors have been used to overexpress IL-10 in donor lungs either prior to (ex vivo) or after orthotopic lung transplant (in vivo), limiting inflammation and allograft rejection.61 Other studies have shown success in transferring genes to decrease disease severity in mouse ALI/ARDS and asthma models, as well as in CF.62–64 Based on these and others preliminary successes, clinical trial of lentiviral gene delivery to treat CF is also in progress.63,65 Despite these many early clinical trials and seeming benefit, the worry of oncogenic gene integration remains. HIV and laboratory lentiviruses have shown a causative effect of oncogenesis in mouse and human. This concern can be reduced by modifications removing replication and integration activities but this also removes a major lentiviral benefit.66–68 Lentivirus-transduced cells are tested for these issues prior to transplant but such testing is not amenable for a direct treatment approach in which the virus is administered directly to the patient. Future modifications to lentivirus may reduce or eliminate these concerns. With successes starting to be shown, lentivirus has entered the clinic for indirect transplant approaches and direct treatment approaches are likely in the near future.
Non-viral vectors
Liposomes
Liposome-mediated transfection, or lipofection, was first developed in the late 1980s and has flourished in vitro with relatively low cytotoxicity compared to a host of other transfection methods. However, usage in vivo has been limited due to interference from serum and lower efficiencies than other types of vectors. Newer liposome formulations have been able to increase efficiency and reduce serum interference allowing better in vivo efficiencies. The greatest advantage, however, of liposomes and other non-viral methods of gene delivery is that they are much less inflammatory and immunogenic than their viral counterparts. Lipofection uses cationic lipid complexes to encapsulate negatively charged cargo yielding a net positively charged complex that can then further interact with the cell’s membrane. Specialized formulations and combinations of lipids may be specific for complexing DNA, RNA, or even nucleoprotein complexes, and can even be specific for the size of the cargo being delivered. These lipopolyplexes are often preferred over other methods due to their relatively low cost, speed, and ease of use. Additionally, at least in vitro, liposome-mediated transfection often has reduced cytotoxicity compared to viruses or electroporation, albeit in a formula-dependent manner, though it most often does not share the same efficiency as these other methods for cargo delivery.69–71 Understanding of the precise mechanisms of lipofection is incomplete, although most lipid complexes are thought to be endocytosed prior to fusion with the endosomal membrane and the resulting cargo release into the cytoplasm.72 Once endocytosed and released into the cytosol, liposomes are dispersed in cells by Brownian motion, avoiding lysosomal or endosomal degradation, thereby allowing them to deliver RNA as well as DNA. However, if they are not unpackaged to release their DNA or RNA as lipid-free species, they do not efficiently form protein-DNA or -RNA complexes for association with microtubule-based motors and fail to efficiently traffic to the nucleus, thereby making nuclear delivery of cargo inefficient. Nucleic acid delivery is limited by the size of the formed liposomes although to a much lesser degree than viral vectors. As a result lipofection can deliver supercoiled plasmid DNA and small RNAs for RNAi, but may have difficulty with large mRNAs.72 The efficiency of lipofection varies by cell type, serum, and antibiotic presence, limiting in vivo capabilities to more topical tissues without further understanding of the necessary formulations. Current lipofection technology is also unable to transfect all cell types, though with better understanding of uptake mechanism and cell membrane characteristics of specific cell types, this may change. Finally, since liposomes carry just the DNA and no associated proteins for recombination, integration of delivered genes occurs extremely infrequently, especially for supercoiled plasmids.
Perhaps the greatest advantage of liposomal transfection in vivo is that fact that most liposomes are much less inflammatory than any viral vectors used. Liposomes have low immunogenicity since there are no protein epitopes to target. Reduced inflammation allows liposomes to be used under already inflammatory conditions to deliver genes with less potential for adverse side-effects that immunogenic (viral) vectors may have. Currently, delivery of anti-inflammatory agents by this approach has focused on delivering traditional pharmaceuticals69–71,73–75 but delivery of nucleic acids, such as for RNAi, is also possible depending on formulation. Low immunogenicity also allows multiple doses to be delivered without alteration of the vector or cargo.76–78 Multiple doses allow for multiple delivery attempts and for a dose-number increase in efficiency that is not possible for viral vectors. Proof of concept multi-dose delivery approaches using liposomes have been performed that successfully tested this hypothesis in both mice and humans.77,79,80
Perhaps the greatest successes for liposome-mediated transfection in the lung have been for delivery of CFTR, giving partial correction of the CFTR defect in mice, sheep, and other models. In most cases, various cationic liposomes or combinations of cationic and other lipid derivatives have been used. While the first use of aerosolized liposomes in the lung delivered a reporter gene in 1992,81 less than one year later CFTR was delivered to the lungs of mice using the same approach using either lipofectin or DC-cholesterol/DOPE,82,83 achieving correction of ion channel defects in at least some animals. Moreover, several clinical trials for CF have been carried out using liposomes and plasmids. The first use of liposomal gene delivery to treat cystic fibrosis occurred in a 1999 double-blind phase 1 trial. The trial showed some short-term benefit but patients in both treatment and placebo (liposome without vector) groups did experience some pulmonary side-effects that spontaneously resolved.76 A more recent phase 2 b trial in 2015 by the same group used a CpG-free plasmid design and a CMV instead of GM-CSF promoter resulting in reduced side-effects and longer term expression.80 Unfortunately for both trials, the results while favorable were minimal due to delivery efficacy.65,76,80 Future work with liposomes in vivo will likely remain limited without further modifications to formulations and plasmid design.
Nanoparticles
Nanoparticles (NPs) are defined as particles on the nanometer scale and, for the purpose of this article, are defined as being 100 nm or less. We further exclude lipid-based formulations (e.g. liposomes), which may otherwise fall under this classification, and which are increasingly being referred to as nanoparticles. NPs have been studied for decades, but only recently have practical applications for their use become feasible. NPs show great promise but have not been as well characterized in comparison to other vectors in terms of vector-host interactions. There are hundreds of thousands of natural or synthetic NPs possible, along with derivatives of each. These derivatives can each have their own chemical/biological properties, making characterization difficult and individualized for given particles. Some common types of NPs include dendrimers, cell penetrant peptides (CPPs), copolymers (e.g. PEGylated poly-l-lysine), and nucleic acid aptamers, among a host of others (for a brief review on clinical NPs, see Bobo et al.84) Dendrimer NPs function similarly to liposomes, forming a spherical structure around the desired cargo before being endocytosed and fusing with the plasma membrane of a cell for delivery. Dendrimer NPs tend to be made of negatively charged amino acids and are often PEGylated in order to aid the formation of the structure.26,85,86 Cell penetrant peptides (CPPs), smaller modified or synthetic peptides, are positively charged, able to associate with negatively charged cargo, such as nucleic acids, and are endocytosed through poorly characterized mechanisms. These CPPs are derivatives of bacterial and viral peptides such as TAT, penetratin, transportan, or oligoarginines, all of which penetrate cellular membranes and can be modified for purposes of gene delivery.70,87 Copolymers like PEGylated poly-l-lysine enter cells by micropinocytosis avoiding endosomes and lysosomes.88 Nucleic acid aptamers can have various properties but are similarly endocytosed for delivery to the cell cytoplasm. A major benefit of some NP designs is a lack of immunogenicity, opening options for multiple repeat doses for treatment of chronic pathology or during inflammatory conditions less conducive to other approaches. Most NPs pass through mucus and surfactant due to their size and general dispersion properties but must be instilled or aspirated as larger droplets in order to reach the deep lung.26,27
Early usage of CPPs for delivery was often somewhat troublesome, as delivered cargo was sequestered in endosomes or had altered function due to complexation with the peptide. Further, synthesis of these peptides was expensive and difficult to do accurately. As a result, other vectors for delivery have been preferentially studied and used. However, commercially available Pepfect and Viromer, as well as the recent development of fmoc solid-phase peptide synthesis has greatly improved the financial feasibility of creation and use of peptide-based nanoparticles.89,90 Trying to overcome the alterations to cargo caused by complexation with the various peptides has resulted in several modifications of design, such as inclusion of redox sensitivity and cyclization of the peptdies.70,91 Some peptides, such as a redox sensitive cyclic amphipathic peptide, have improved efficiency of RNAi delivery to the lung compared to both Pepfect and lipofection.70 However, CPPs and these cyclic peptide derivatives have not been used to deliver larger plasmid DNAs, a primary approach to gene delivery. While other larger polymer-based NPs such as dendrimers, nucleic acid aptamers, and copolymers, can deliver plasmids with varying efficiencies depending on formulation and target cells, most NPs seem to show greater abilities for smaller RNAi delivery applications. Some NPs including CPPs can also deliver small proteins, opening a potential use over viral vectors for CRISPR-related technologies. One concern of NPs is a buildup of the particles in cells or in filtering organs, such as the kidney or liver, and work still needs to be done to characterize biodistribution of nanoparticles used for gene delivery.26,70,85,89,92 Despite these possible concerns, a number of NPs have been used for in vivo gene delivery to the lung with some early success, including for delivery of plasmids up to 20 kb93 and siRNA70 in mice. Studies using PEGylated poly-l-lysine copolymers that form <20 nm particles, copernicus therapeutics has also shown safety and short-term efficacy of delivered plasmids in mice and in CF patients.94–96 Unfortunately, long-term expression of delivered plasmids was not obtained though this could potentially be overcome by altering plasmid design or by using other cargo. With a better understanding of uptake, delivery, and biodistribution characteristics, NPs will likely enter the clinic for anti-inflammatory or multi-dose uses.
Electroporation
Electroporation (EP) was first used in vivo in the late 1980s and is currently the only physical method that is practical or feasible for use in the lung.97 In vivo EP for the purpose of gene delivery, or gene electro transfer (GET) has been used sparingly, but electrochemotherapy (ECT; electroporation at similar fields to gene delivery but for chemotherapeutic drug delivery) and irreversible electroporation (IRE; electroporation at much higher field strengths than for gene transfer) have been used for multiple cancer treatment approaches.98,99 In research settings, EP is a staple laboratory technique for in vitro gene delivery, especially in hard-to-transfect cell lines. EP creates transient pores in the plasma membrane due to charge distribution across the membrane and electrochemical rearrangement of membrane phospholipids into transient non-selective pores. Pores from electroporation vary in number and size (∼1–400 nm) depending on cell membrane characteristics, voltage, duration, and type of electric pulse.100,101 Nucleic acids and other material in the extracellular solution can then enter the cell via these pores. Following removal of the electric field, the pores close through endogenous membrane repair mechanisms. One of the prime benefits of using a physical method of gene delivery is that DNA, RNA, or other cargo can be delivered with no carrier (e.g. complexed lipid or viral proteins) to alter cellular distribution or cargo activity. Exogenous DNA from viruses or bacteria does not enter a cell unless purposefully included as cargo themselves. This feature has become of great use for ribonucleoprotein CRISPR complexes to deliver functional CRISPR-Cas while bypassing the need for exogenous components.102–104
EP is highly adaptable; the voltage, pulse, and other characteristics can all be altered to fit the cell type being transfected, as well as the amount and sizes of the pores. Because EP physically affects the phospholipids that are components of all cells, it can be used on all cell types with appropriate empirical changes to the voltage, pulse duration, and other variables. Unfortunately, in vitro EP often has an unfavorable effect on cell viability, with upwards of 30–40% cytotoxicity. However, in vivo cellular and tissue recovery from GET are much more robust with no significant harm under appropriate settings.15,105,106 Usage of EP in IRE and ECT has shown that EP is safe to use in vivo in humans even for highly vascularized tissue such as liver, pancreas, and lung.107–110 Clinical use of GET has generally been applied to the skin, as simple electrical pads and an adjustable power supply are all that is required. This approach has been highly successful for overexpression of immune regulators to treat melanoma or for DNA-based vaccine delivery.111–114 For delivery to internal organs, such as the lung, we have achieved success using simple defibrillation pads on the chest in mice, rats, and pigs.15,105,106,115 Once purified, plasmid DNA is administered to the lungs by aspiration, intratracheal injection, nebulization, or aerosolization, electric pulses are delivered across the chest. This approach has been used successfully for gene transfer of reporter genes in mice,116 rats,117 pigs,118 and sheep,119 as well as a number of therapeutic genes to treat ALI/ARDS,15,117,120–125 pulmonary fibrosis,126–128 and surfactant protein B deficiency.129 Following electroporation, plasmid DNA is transferred to and expresses in all cell types throughout the lung, including surface airway and alveolar epithelial cells as well as cells underneath the epithelium, including interstitial fibroblasts, airway and vascular smooth muscle cells, and endothelial cells. By manipulating the promoter or other sequence elements on the plasmid, this means that targeted gene transfer to specific cell types can be achieved.128,130,131
The current clinical focus for EP is mainly on IRE, a method for tissue ablation using targeted high energy EP, and ECT, using EP to target cancerous tissue with normally cell-impermeant chemotherapeutic drugs.107,108,132,133 IRE applies > 2000 V/cm fields with multiple pulses up to 30 ms each, while ECT and GET typically use 250–500 V/cm for 1 ms or less to instead deliver chemotherapeutic drugs and genes, respectively.107–110,113,132–135 At all of these field strengths, EP has proven safe in animal models and humans. In human-sized 50 kg pigs, delivery of EP pulses across the chest for lung gene delivery uses less than 4 J of energy for gene transfer. By comparison, an automated external defibrillator (AED) uses 360 J in a single 1 ms pulse to defibrillate a patient and is deemed safe. Perception of GET as being related to electroshock therapy or electrocution, and the confusion with IRE and its use as an ablative therapy, has hindered its translation to the clinic. However, as safety has been clearly demonstrated and with more research on the necessary delivery characteristics, EP for gene delivery to the lung will likely translate to the clinic in the future as well.
Future uses
Viruses
Viral vectors, such as lentivirus, are still some of the best vectors for long-term expression of transgenes with genome integration. Additionally, co-infection of cells and tissues with constructs containing transposons can allow for integration of transgenes carried by other viruses, including adenovirus, thus overcoming the limitation of this virus for only short duration expression. Such transposons can also be used in conjunction with non-viral gene delivery approaches. An example of a transposon, piggybac, functions to integrate delivered genes into the host genome based on the transposase recognition of short repeat sequences.8 Transposon-based integration allows for long-term maintained expression, but there are multiple integration sites due to the short recognition sequence of the transposase, and possible off-target effects have not fully been elucidated. Transposons are not limited to combined use with viral vectors. Indeed, delivery of plasmid-based transposons to the lung has also been achieved using polymers, nanoparticles, and electroporation to achieve integration and long-term expression.136,137 Modified viral vectors for in vivo gene delivery also are being further developed for vaccine creation with the benefit that they induce mild local and systemic inflammation which can aid in generating an immune response. Viruses are used in many of the FDA-approved cellular/gene therapies where cells are removed, modified, and reinjected, but this type of ex vivo to in vivo usage is not suited for scenarios of acute disease where time may be of the essence or with other difficulties of transplantation. Viral targeting is often not unique to specific cells or tissues as receptors utilized for viral uptake are often widespread. Viral vectors are also limited by their genome size and the amount of DNA/RNA that they can package in their capsids, thus reducing their capacity to deliver large genes or multiple genes. However, even with these limitations, viral vectors are an established means of gene transfer that will likely have a long future of clinical use.
Lipofection
Despite being the major “go-to” technology for cellular transfection in the laboratory, in vivo lipofection will likely remain a secondary strategy unless its efficiency in the living animal can be increased. One long perceived limitation of non-viral methods has been a lack of long-term expression. This is not the case, since non-viral based plasmids with the appropriate promoter choice are suitable for long-term expression, but without any side effects caused by viral vectors. These non-integrated plasmids delivered by a host of non-viral methods have shown long-term expression in a variety of tissues in mice, with reports of expression six months to one year after delivery to mice.128,138–140 In conjunction with other technology such as electroporation or nanoparticles, lipofection may have greater potential than its use alone. Further improvements in lipid chemistry and structure may also increase their use in vivo. A major advantage is that liposomes have essentially no size constraints for cargo, allowing delivery of large genes and also have a relatively low inflammatory profile.
Nanoparticles
NPs have only recently become practical to use, and recent data has attempted to explain how NPs work. Mechanisms of entry can be different between cell types and NPs making “full” understanding of all particle types an ongoing effort. Functional understanding of how each specific NP is internalized will lead to better structural formulations for improved targeting and uptake. There are many currently untested NP formulations that will allow us to further understand how NPs “capture” cargo, are endocytosed, and allow the cargo to remain functional. NPs are highly versatile, can have low immunogenicity, low cytotoxicity, lack exogenous DNA or RNA (unlike exosomes), and may have higher efficiency compared to viral vectors, lipofection, or EP methods in certain circumstances. NPs are able to deliver large genes and plasmids depending on formulation. As yet, NPs have not been shown to induce long-term expression of delivered plasmids and are therefore a short-term delivery approach without additional factors or appropriate promoter choice. Repeated dosing approaches may increase their versatility but may also generate side-effects, and biodistribution and clearance will be important to this use. NPs as drug and gene delivery mechanisms are likely to progress relatively slowly as they remain more expensive than more classical approaches but may gain usage as production becomes cheaper and mechanistic understanding is improved. Clinically, since they are seen as very similar to classical pharmacologic drugs, their acceptance as “safe”, along with supporting safety data, may greatly aid their acceptance.
Electroporation
Electroporation is likely to be developed more quickly for both tumor ablation (IRE) and chemotherapy (ECT), improving parameters for efficiency and targeting. EP has already entered the clinic for these anti-tumor uses and will likely remain prior to the development of more personalized therapies. IRE and ECT showcase EP as safe, therefore in vivo EP for gene delivery may see a rise in research and clinical usage. Electroporation for gene delivery will also certainly be further developed both as a stand-alone delivery approach for classic overexpression of plasmid DNA and when complexed with other technology, such as CRISPR. EP can deliver large genes and multiple genes or components in a single delivery event. EP is targeted primarily through the placement of electrodes and parameters used. Future possibilities of EP would see further development of research technologies that improve targeting of gene delivery to specific organs or cell types. EP could also potentially be used in combination with nanoparticles or liposomes for nuclear or organelle specific delivery. Use of EP with integration factors or long-term expressing plasmids (CpG reduced) is also in progress.113,141 Due to the lack of any exogenous components (i.e. lipid, polymers, other carriers) other than the nucleic acid that are needed for delivery, electroporation is very likely to attain FDA-approval for gene therapy. Indeed, there are currently over 95 clinical trials for electroporation-mediated DNA gene transfer underway, including a Phase II trial for a SARS-CoV2 vaccine against Covid-19, attesting to its clinical appeal.
Exosomes
Exosomes were first discovered in 1983 as microvesicles less than 150 nm that are released from cells. They are endogenous cell-membrane derived vesicles that have been shown to contain RNAs, proteins, and other molecules involved in cell–cell signaling.23,142 Because exosomes are composed of normal cellular membranes, they enter cells by innate clathrin-dependent and macropinocytic mechanisms after ligand-receptor or membrane interactions.143 Further, since they are seen as “self”, they largely avoid initiating an inflammatory response. Some tissue-specific targeting of exosomes dependent on integrins expressed by the exosomes has been shown but its mechanisms are not completely understood.23 The main limitation for their use comes from challenges in their isolation and purification at scales needed for commercial or human use. Separating exosomes containing the correct factors from other exosomes and cell debris requires density gradients or ultracentrifugation, and/or other filtration techniques, which can vary in terms of what specific exosomes are isolated. Further, keeping exosomes stable long-term has proven difficult, as they are prone to aggregate and degrade.144–149 Additionally, methods for loading cargo into exosomes currently rely on nanoparticles or electroporation-based delivery to the exosome.144 A potential drawback is that cellular components internal to cellularly-produced exosomes may also be delivered along with any loaded cargo.142,150–152 Exosomes have been used in mice for delivery to brain144and lymph nodes151 without noticeable toxicity as long as the cell source was syngeneic. Safety studies will be the first major hurdle for exosome delivery to humans since unapproved exosome containing products (from the cells in which the exosomes are produced) have been warned against by the FDA.153 However, recent methods for exosome production from red blood cells may greatly reduce some of these issues.154
Exosomes are farthest from clinic but likely to increase in usage due to favorable characteristics of a cellular origin and endogenous uptake. The development of technology allowing greater specificity in the separation of specific exosome subsets from other subsets and other vesicles will be important to understanding and clinical usage. Further development of a solution/method that is better at maintaining exosome stability will also be important to any widespread clinical usage and pharmaceutical feasibility. Small particle flow cytometers are now being deployed in many research cores and will help further understand exosome characteristics in conjunction with proteomics and genomics. Understanding of exosome characteristics will allow better tissue specificity, and possible generation of synthetic exosomes. No cargo size limit is known for exosomes, but it will likely be difficult to load large plasmids without compromising exosome membrane integrity. Exosomes can deliver multiple plasmids in multiple doses but as a single dose would be limited by “loading” of exosomes with plasmids and the uptake of target cells. Exosome delivered RNAi or plasmid expression have not been studied long-term, with 48 h or less timepoints being common.144,151,155 The duration of delivered nucleic acid(s) persistence in tissues will be important to clinical usage. Improvement of storage media and conditions may help prolong exosome stability and increase pharmacologic capabilities.
Conclusions
Despite drawbacks of the reviewed methods of gene delivery to lung, they are currently the most prominently used in research and clinical trials. Several viruses and electroporation gene delivery uses are already approved or being tested in clinical trials. Liposomes will be limited in use and used primarily for non-integrative treatment approaches. Nanoparticles will likely overtake liposome use by recapitulating necessary characteristics along with other modification. Nanoparticles will likely be developed for other uses, finding niches where lipofection falls short as they enter clinical trials. Exosome use is novel and while favorable due to endogenous characteristics is dependent on further development of associated technology. Better comprehension of necessary characteristics has the potential to develop exosomes into targeted treatments. Understanding these gene delivery techniques from both a basic science and clinical perspective is necessary to ensure proper patient care and further development of the technologies.
AUTHORS’ CONTRIBUTIONS
All authors participated in the interpretation of the studies and review of the manuscript. UKB wrote the majority of the manuscript and DAD revised the manuscript, providing additional sources, and guidance.
ACKNOWLEDGMENTS
The authors would like to thank Kaitlyn Shaw for artwork in Figure 1.
Footnotes
Declaration OF CONFLICTING INTERESTS: The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.
Funding: This work was funded in part from grants HL138538, HL148825, HL131143, and HL120521 from the National Institutes of Health and from a research grant from the Cystic Fibrosis Foundation.
ORCID iDs
Uday K Baliga https://orcid.org/0000-0002-4557-8645
David A Dean https://orcid.org/0000-0001-7748-8073
References
- 1.de Miguel-Beriain I. The ethics of stem cells revisited. Adv Drug Deliv Rev 2015; 82-83:176–80 [DOI] [PubMed] [Google Scholar]
- 2.Manzar N, Manzar B, Hussain N, Hussain MFA, Raza S. The ethical dilemma of embryonic stem cell research. Sci Eng Ethics 2013; 19:97–106 [DOI] [PubMed] [Google Scholar]
- 3.Caplan A, Purves D. A quiet revolution in organ transplant ethics. J Med Ethics 2017; 43:797–800 [DOI] [PubMed] [Google Scholar]
- 4.Trey T, Caplan AL, Lavee J. Transplant ethics under scrutiny - responsibilities of all medical professionals. Croat Med J 2013; 54:71–4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Gene therapy’s next installment. Nat Biotechnol 2019; 37:697. [DOI] [PubMed] [Google Scholar]
- 6.Normile D. iPS cell therapy reported safe. Science 2017; 355:1109–10 [DOI] [PubMed] [Google Scholar]
- 7.Salzman R, Cook F, Hunt T, Malech HL, Reilly P, Foss-Campbell B, Barrett D. Addressing the value of gene therapy and enhancing patient access to transformative treatments. Mol Ther 2018; 26:2717–26 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Zheng Y, Li Z-R, Yue R, Fu Y-L, Liu Z-Y, Feng H-Y, Li J-G, Han S-Y.. PiggyBac Transposon System with Polymeric Gene Carrier Transfected into Human T Cells. Am J Transl Res 2019; 11: 7126–36 [PMC free article] [PubMed] [Google Scholar]
- 9.Bofill-De Ros X, Gu S. Guidelines for the optimal design of miRNA-based shRNAs. Methods 2016; 103:157–66 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Rao DD, Vorhies JS, Senzer N, Nemunaitis J. siRNA vs. shRNA: similarities and differences. Adv Drug Deliv Rev 2009; 61:746–59 [DOI] [PubMed] [Google Scholar]
- 11.Wang Q, Vossen A, Ikeda Y, Devaux P. Measles vector as a multigene delivery platform facilitating iPSC reprogramming. Gene Ther 2019; 26:151–64 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Rafeeq MM, Murad HAS. Cystic fibrosis: current therapeutic targets and future approaches. J Transl Med 2017; 15:84–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Villanueva G, Marceniuk G, Murphy MS, Walshaw M, Cosulich R. Diagnosis and management of cystic fibrosis: summary of NICE guidance. BMJ 2017; 359:j4574–j74 [DOI] [PubMed] [Google Scholar]
- 14.Dickson RP, Singer BH, Newstead MW, Falkowski NR, Erb-Downward JR, Standiford TJ, Huffnagle GB. Enrichment of the lung microbiome with gut bacteria in sepsis and the acute respiratory distress syndrome. Nat Microbiol 2016; 1:16113–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Emr BM, Roy S, Kollisch-Singule M, Gatto LA, Barravecchia M, Lin X, Young JL, Wang G, Liu J, Satalin J, Snyder K, Nieman GF, Dean DA. Electroporation-mediated gene delivery of Na+,K+ -ATPase, and ENaC subunits to the lung attenuates acute respiratory distress syndrome in a two-hit porcine model. Shock 2015; 43:16–23 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Zemanick ET, Accurso FJ. Entering the era of highly effective CFTR modulator therapy. Lancet 2019; 394:1886–88 [DOI] [PubMed] [Google Scholar]
- 17.Chiuchiolo MJ, Crystal RG. Gene therapy for alpha-1 antitrypsin deficiency lung disease. Ann Am Thorac Soc 2016; 13(Suppl 4):S352–S69 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Edgar RG, Patel M, Bayliss S, Crossley D, Sapey E, Turner AM. Treatment of lung disease in alpha-1 antitrypsin deficiency: a systematic review. Int J Chron Obstruct Pulmon Dis 2017; 12:1295–308 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Hammarsten JF, Welch MH, Richardson RH, Patterson CD, Guenter CA. Familial alpha-1, antitrypsin deficiency and pulmonary emphysema. Transac Am Clin Climatol Assoc 1969; 80:7–14 [PMC free article] [PubMed] [Google Scholar]
- 20.Hurst A. Familial emphysema. Am Rev Respir Dis 1959; 80:179–80 [DOI] [PubMed] [Google Scholar]
- 21.Meyer KC. Pulmonary fibrosis, part I: epidemiology, pathogenesis, and diagnosis. Exp Rev Respir Med 2017; 11:343–59 [DOI] [PubMed] [Google Scholar]
- 22.Thannickal VJ, Toews GB, White ES, Lynch Iii JP, Martinez FJ. Mechanisms of pulmonary fibrosis. Ann Rev Med 2004; 55:395–417 [DOI] [PubMed] [Google Scholar]
- 23.Hoshino A, Costa-Silva B, Shen T-L, Rodrigues G, Hashimoto A, Tesic Mark M, Molina H, Kohsaka S, Di Giannatale A, Ceder S, Singh S, Williams C, Soplop N, Uryu K, Pharmer L, King T, Bojmar L, Davies AE, Ararso Y, Zhang T, Zhang H, Hernandez J, Weiss JM, Dumont-Cole VD, Kramer K, Wexler LH, Narendran A, Schwartz GK, Healey JH, Sandstrom P, Labori KJ, Kure EH, Grandgenett PM, Hollingsworth MA, de Sousa M, Kaur S, Jain M, Mallya K, Batra SK, Jarnagin WR, Brady MS, Fodstad O, Muller V, Pantel K, Minn AJ, Bissell MJ, Garcia BA, Kang Y, Rajasekhar VK, Ghajar CM, Matei I, Peinado H, Bromberg J, Lyden D. Tumour exosome integrins determine organotropic metastasis. Nature 2015; 527:329–35 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Maru Y. The lung metastatic niche. J Mol Med 2015; 93:1185–92 [DOI] [PubMed] [Google Scholar]
- 25.Fu J, Fiegel J, Krauland E, Hanes J. New polymeric carriers for controlled drug delivery following inhalation or injection. Biomaterials 2002; 23:4425–33 [DOI] [PubMed] [Google Scholar]
- 26.Thorley AJ, Ruenraroengsak P, Potter TE, Tetley TD. Critical determinants of uptake and translocation of nanoparticles by the human pulmonary alveolar epithelium. ACS Nano 2014; 8:11778–89 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Driscoll KE, Costa DL, Hatch G, Henderson R, Oberdorster G, Salem H, Schlesinger RB. Intratracheal instillation as an exposure technique for the evaluation of respiratory tract toxicity: uses and limitations. Toxicol Sci 2000; 55:24–35 [DOI] [PubMed] [Google Scholar]
- 28.Bett AJ, Haddara W, Prevec L, Graham FL. An efficient and flexible system for construction of adenovirus vectors with insertions or deletions in early regions 1 and 3. Proc Natl Acad Sci U S A 1994; 91:8802–6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Hendrickx R, Stichling N, Koelen J, Kuryk L, Lipiec A, Greber UF. Innate immunity to adenovirus. Hum Gene Ther 2014; 25:265–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Nayak S, Herzog RW. Progress and prospects: immune responses to viral vectors. Gene Ther 2010; 17:295–304 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Yang Y, Nunes FA, Berencsi K, Furth EE, Gönczöl E, Wilson JM. Cellular immunity to viral antigens limits E1-deleted adenoviruses for gene therapy. Proc Natl Acad Sci U S A 1994; 91:4407–11 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Ng Nb-P P. Progress towards liver and lung-directed gene therapy with helper- dependent adenoviral vectors. Curr Gene Ther 2009; 9:329–40 [DOI] [PubMed] [Google Scholar]
- 33.van Haasteren J, Hyde SC, Gill DR. Lessons learned from lung and liver in-vivo gene therapy: implications for the future. Exp Opin Biol Ther 2018; 18:959–72 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Aggarwal C, Haas AR, Metzger S, Aguilar LK, Aguilar-Cordova E, Manzanera AG, Gómez-Hernández G, Katz SI, Alley EW, Evans TL, Bauml JM, Cohen RB, Langer CJ, Albelda SM, Sterman DH. Phase I study of intrapleural gene-mediated cytotoxic immunotherapy in patients with malignant pleural effusion. Mol Ther 2018; 26:1198–205 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Kim JW, Glasgow JN, Nakayama M, Ak F, Ugai H, Curiel DT. An adenovirus vector incorporating carbohydrate binding domains utilizes glycans for gene transfer. PLoS One 2013; 8:e55533–e33 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Lenman A, Liaci AM, Liu Y, Frängsmyr L, Frank M, Blaum BS, Chai W, Podgorski II, Harrach B, Benkő M, Feizi T, Stehle T, Arnberg N. Polysialic acid is a cellular receptor for human adenovirus 52. Proc Natl Acad Sci U S A 2018; 115:E4264–E73 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Sterman DH, Recio A, Carroll RG, Gillespie CT, Haas A, Vachani A, Kapoor V, Sun J, Hodinka R, Brown JL, Corbley MJ, Parr M, Ho M, Pastan I, Machuzak M, Benedict W, Zhang X-Q, Lord EM, Litzky LA, Heitjan DF, June CH, Kaiser LR, Vonderheide RH, Albelda SM. A phase I clinical trial of single-dose intrapleural IFN-β gene transfer for malignant pleural mesothelioma and metastatic pleural effusions: high rate of antitumor immune responses. Clin Cancer Res 2007; 13:4456–66 [DOI] [PubMed] [Google Scholar]
- 38.Gomez-Gutierrez JG, Nitz J, Sharma R, Wechman SL, Riedinger E, Martinez-Jaramillo E, Sam Zhou H, McMasters KM. Combined therapy of oncolytic adenovirus and temozolomide enhances lung cancer virotherapy in vitro and in vivo. Virology 2016; 487:249–59 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Lino CA, Harper JC, Carney JP, Timlin JA. Delivering CRISPR: a review of the challenges and approaches. Drug Deliv 2018; 25:1234–57 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Cooney AL, Singh BK, Loza LM, Thornell IM, Hippee CE, Powers LS, Ostedgaard LS, Meyerholz DK, Wohlford-Lenane C, Stoltz DA, B McCray P, Jr, Sinn PL. Widespread airway distribution and short-term phenotypic correction of cystic fibrosis pigs following aerosol delivery of piggyBac/adenovirus. Nucl Acids Res 2018; 46:9591–600 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Grieger JC, Samulski RJ. Packaging capacity of adeno-associated virus serotypes: impact of larger genomes on infectivity and postentry steps. J Virol 2005; 79:9933–44 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Wu Z, Yang H, Colosi P. Effect of genome size on AAV vector packaging. Molecular therapy: the journal of the american society of. Gene Ther 2010; 18:80–6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Naso MF, Tomkowicz B, Perry 3rd WL, Strohl WR. Adeno-associated virus (AAV) as a vector for gene therapy. BioDrugs 2017; 31:317–34 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Verdera HC, Kuranda K, Mingozzi F. AAV vector immunogenicity in humans: a long journey to successful gene transfer. Mol Ther 2020; 28:723–46 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Guggino WB, Cebotaru L. Adeno-associated virus (AAV) gene therapy for cystic fibrosis: current barriers and recent developments. Exp Opin Biol Ther 2017; 17:1265–73 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Steines B, Dickey DD, Bergen J, Excoffon KJ, Weinstein JR, Li X, Yan Z, Abou Alaiwa MH, Shah VS, Bouzek DC, Powers LS, Gansemer ND, Ostedgaard LS, Engelhardt JF, Stoltz DA, Welsh MJ, Sinn PL, Schaffer DV, Zabner J. CFTR gene transfer with AAV improves early cystic fibrosis pig phenotypes. JCI Insight 2016; 1:e88728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Duncan GA, Kim N, Colon-Cortes Y, Rodriguez J, Mazur M, Birket SE, Rowe SM, West NE, Livraghi-Butrico A, Boucher RC, Hanes J, Aslanidi G, Suk JS. An adeno-associated viral vector capable of penetrating the mucus barrier to inhaled gene therapy. Mol Ther 2018; 9:296–304 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Al-Zaidy SA, Mendell JR. From clinical trials to clinical practice: practical considerations for gene replacement therapy in SMA type 1. Pediatr Neurol 2019; 100:3–11 [DOI] [PubMed] [Google Scholar]
- 49.Darrow JJ. Luxturna: FDA documents reveal the value of a costly gene therapy. Drug Discov Today 2019; 24:949–54 [DOI] [PubMed] [Google Scholar]
- 50.Hoy SM. Onasemnogene abeparvovec: first global approval. Drugs 2019; 79:1255–62 [DOI] [PubMed] [Google Scholar]
- 51.Patel U, Boucher M, de Leseleuc L, Visintini S. Voretigene neparvovec: an emerging gene therapy for the treatment of inherited blindness CADTH Issues in Emerging Health Technologies, 2016, p. 169– [PubMed]
- 52.Russell S, Bennett J, Wellman JA, Chung DC, Yu Z-F, Tillman A, Wittes J, Pappas J, Elci O, McCague S, Cross D, Marshall KA, Walshire J, Kehoe TL, Reichert H, Davis M, Raffini L, George LA, Hudson FP, Dingfield L, Zhu X, Haller JA, Sohn EH, Mahajan VB, Pfeifer W, Weckmann M, Johnson C, Gewaily D, Drack A, Stone E, Wachtel K, Simonelli F, Leroy BP, Wright JF, High KA, Maguire AM. Efficacy and safety of voretigene neparvovec (AAV2-hRPE65v2) in patients with RPE65-mediated inherited retinal dystrophy: a randomised, controlled, open-label, phase 3 trial. Lancet 2017; 390:849–60 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Counsell JR, Asgarian Z, Meng J, Ferrer V, Vink CA, Howe SJ, Waddington SN, Thrasher AJ, Muntoni F, Morgan JE, Danos O. Lentiviral vectors can be used for full-length dystrophin gene therapy. Sci Rep 2017; 7:79–79 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Usme-Ciro JA, Campillo-Pedroza N, Almazán F, Gallego-Gomez JC. Cytoplasmic RNA viruses as potential vehicles for the delivery of therapeutic small RNAs. Virol J 2013; 10:185–85 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Rossetti M, Cavarelli M, Gregori S, Scarlatti G. HIV-derived vectors for gene therapy targeting dendritic cells . In: Wu L, Schwartz O (eds) HIV interactions with dendritic cells: infection and immunity. New York, NY: Springer, 2013, pp. 239–61 [DOI] [PubMed]
- 56.Schlimgen R, Howard J, Wooley D, Thompson M, Baden LR, Yang OO, Christiani DC, Mostoslavsky G, Diamond DV, Duane EG, Byers K, Winters T, Gelfand JA, Fujimoto G, Hudson TW, Vyas JM. Risks associated with lentiviral vector exposures and prevention strategies. J Occup Environ Med 2016; 58:1159–66 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Biffi A, Montini E, Lorioli L, Cesani M, Fumagalli F, Plati T, Baldoli C, Martino S, Calabria A, Canale S, Benedicenti F, Vallanti G, Biasco L, Leo S, Kabbara N, Zanetti G, Rizzo WB, Mehta NAL, Cicalese MP, Casiraghi M, Boelens JJ, Del Carro U, Dow DJ, Schmidt M, Assanelli A, Neduva V, Di Serio C, Stupka E, Gardner J, von Kalle C, Bordignon C, Ciceri F, Rovelli A, Roncarolo MG, Aiuti A, Sessa M, Naldini L. Lentiviral hematopoietic stem cell gene therapy benefits metachromatic leukodystrophy. Science 2013; 341:1233158–58 [DOI] [PubMed] [Google Scholar]
- 58.Sessa M, Lorioli L, Fumagalli F, Acquati S, Redaelli D, Baldoli C, Canale S, Lopez ID, Morena F, Calabria A, Fiori R, Silvani P, Rancoita PMV, Gabaldo M, Benedicenti F, Antonioli G, Assanelli A, Cicalese MP, del Carro U, Sora MGN, Martino S, Quattrini A, Montini E, Di Serio C, Ciceri F, Roncarolo MG, Aiuti A, Naldini L, Biffi A. Lentiviral haemopoietic stem-cell gene therapy in early-onset metachromatic leukodystrophy: an ad-hoc analysis of a non-randomised, open-label, phase 1/2 trial. Lancet 2016; 388:476–87 [DOI] [PubMed] [Google Scholar]
- 59.Cavazzana-Calvo M, Payen E, Negre O, Wang G, Hehir K, Fusil F, Down J, Denaro M, Brady T, Westerman K, Cavallesco R, Gillet-Legrand B, Caccavelli L, Sgarra R, Maouche-Chrétien L, Bernaudin F, Girot R, Dorazio R, Mulder G-J, Polack A, Bank A, Soulier J, Larghero J, Kabbara N, Dalle B, Gourmel B, Socie G, Chrétien S, Cartier N, Aubourg P, Fischer A, Cornetta K, Galacteros F, Beuzard Y, Gluckman E, Bushman F, Hacein-Bey-Abina S, Leboulch P. Transfusion independence and HMGA2 activation after gene therapy of human β-thalassaemia. Nature 2010; 467:318–22 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Paul-Smith MC, Pytel KM, Gelinas J-F, McIntosh J, Pringle I, Davies L, Chan M, Meng C, Bell R, Cammack L, Moran C, Cameron L, Inoue M, Tsugumine S, Hironaka T, Gill DR, Hyde SC, Nathwani A, Alton EWFW, Griesenbach U. The murine lung as a factory to produce secreted intrapulmonary and circulatory proteins. Gene Ther 2018; 25:345–58 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Oishi H, Juvet SC, Martinu T, Sato M, Medin JA, Liu M, Keshavjee S. A novel combined ex vivo and in vivo lentiviral interleukin-10 gene delivery strategy at the time of transplantation decreases chronic lung allograft rejection in mice. J Thorac Cardiovasc Surg 2018; 156:1305–15 [DOI] [PubMed] [Google Scholar]
- 62.Cheng S. Lentiviral vector-mediated delivery of lysophosphatidylcholine acyltransferase 1 attenuates airway inflammation in ovalbumin-induced allergic asthmatic mice. Asian Pacific J Allergy Immunol 2015; 33:320–29 [DOI] [PubMed] [Google Scholar]
- 63.Marquez Loza LI, Yuen EC, McCray Jr PB. Lentiviral vectors for the treatment and prevention of cystic fibrosis lung disease. Genes 2019; 10:218–18 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Sun G-Y, Yang H-H, Guan X-X, Zhong W-J, Liu Y-P, Du M-Y, Luo X-Q, Zhou Y, Guan C-X. Vasoactive intestinal peptide overexpression mediated by lentivirus attenuates lipopolysaccharide-induced acute lung injury in mice by inhibiting inflammation. Mol Immunol 2018; 97:8–15 [DOI] [PubMed] [Google Scholar]
- 65.Alton EWFW, Beekman JM, Boyd AC, Brand J, Carlon MS, Connolly MM, Chan M, Conlon S, Davidson HE, Davies JC, Davies LA, Dekkers JF, Doherty A, Gea-Sorli S, Gill DR, Griesenbach U, Hasegawa M, Higgins TE, Hironaka T, Hyndman L, McLachlan G, Inoue M, Hyde SC, Innes JA, Maher TM, Moran C, Meng C, Paul-Smith MC, Pringle IA, Pytel KM, Rodriguez-Martinez A, Schmidt AC, Stevenson BJ, Sumner-Jones SG, Toshner R, Tsugumine S, Wasowicz MW, Zhu J. Preparation for a first-in-man lentivirus trial in patients with cystic fibrosis. Thorax 2017; 72:137–47 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Mapekula L, Ramorola BR, Goolam Hoosen T, Mowla S. The interplay between viruses & host microRNAs in cancer – an emerging role for HIV in oncogenesis. Crit Rev Oncol Hematol 2019; 137:108–14 [DOI] [PubMed] [Google Scholar]
- 67.Marcucci KT, Jadlowsky JK, Hwang W-T, Suhoski-Davis M, Gonzalez VE, Kulikovskaya I, Gupta M, Lacey SF, Plesa G, Chew A, Melenhorst JJ, Levine BL, June CH. Retroviral and lentiviral safety analysis of gene-modified T cell products and infused HIV and oncology patients. Mol Ther 2018; 26:269–79 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Themis M, Waddington SN, Schmidt M, von Kalle C, Wang Y, Al-Allaf F, Gregory LG, Nivsarkar M, Themis M, Holder MV, Buckley SMK, Dighe N, Ruthe AT, Mistry A, Bigger B, Rahim A, Nguyen TH, Trono D, Thrasher AJ, Coutelle C. Oncogenesis following delivery of a nonprimate lentiviral gene therapy vector to fetal and neonatal mice. Mol Ther 2005; 12:763–71 [DOI] [PubMed] [Google Scholar]
- 69.Salimzadeh L, Jaberipour M, Hosseini A, Ghaderi A. Non-viral transfection methods optimized for gene delivery to a lung cancer cell line. Avicenna J Med Biotechnol 2013; 5:68–77 [PMC free article] [PubMed] [Google Scholar]
- 70.Welch JJ, Swanekamp RJ, King C, Dean DA, Nilsson BL. Functional delivery of siRNA by disulfide-constrained cyclic amphipathic peptides. ACS Med Chem Lett 2016; 7:584–89 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Yanagihara K, Cheng P-W. Lectin enhancement of the lipofection efficiency in human lung carcinoma cells. Biochim Biophys Acta 1999; 1472:25–33 [DOI] [PubMed] [Google Scholar]
- 72.Cardarelli F, Digiacomo L, Marchini C, Amici A, Salomone F, Fiume G, Rossetta A, Gratton E, Pozzi D, Caracciolo G. The intracellular trafficking mechanism of lipofectamine-based transfection reagents and its implication for gene delivery. Sci Rep 2016; 6:25879–79 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Anderson R, Franch A, Castell M, Perez-Cano FJ, Bräuer R, Pohlers D, Gajda M, Siskos AP, Katsila T, Tamvakopoulos C, Rauchhaus U, Panzner S, Kinne RW. Liposomal encapsulation enhances and prolongs the anti-inflammatory effects of water-soluble dexamethasone phosphate in experimental adjuvant arthritis. Arthritis Res Ther 2010; 12:R147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Axelsson B. Liposomes as carriers for anti-inflammatory agents. Adv Drug Deliv Rev 1989; 3:391–404 [Google Scholar]
- 75.Crielaard BJ, Lammers T, Morgan ME, Chaabane L, Carboni S, Greco B, Zaratin P, Kraneveld AD, Storm G. Macrophages and liposomes in inflammatory disease: friends or foes? Int J Pharm 2011; 416:499–506 [DOI] [PubMed] [Google Scholar]
- 76.Alton E, Stern M, Farley R, Jaffe A, Chadwick SL, Phillips J, Davies J, Smith SN, Browning J, Davies MG, Hodson ME, Durham SR, Li D, Jeffery PK, Scallan M, Balfour R, Eastman SJ, Cheng SH, Smith AE, Meeker D, Geddes DM. Cationic lipid-mediated CFTR gene transfer to the lungs and nose of patients with cystic fibrosis: a double-blind placebo-controlled trial. Lancet 1999; 353:947–54 [DOI] [PubMed] [Google Scholar]
- 77.Alton Ewfw Boyd AC, Cheng SH, Davies JC, Davies LA, Dayan A, Gill DR, Griesenbach U, Higgins T, Hyde SC, Innes JA, McLachlan G, Porteous D, Pringle I, Scheule RK, Sumner-Jones S. Toxicology study assessing efficacy and safety of repeated administration of lipid/DNA complexes to mouse lung. Gene Ther 2014; 21:89–95 [DOI] [PubMed] [Google Scholar]
- 78.Stern M, Ulrich K, Geddes DM, Alton EWFW. Poly (D, L-lactide-co-glycolide)/DNA microspheres to facilitate prolonged transgene expression in airway epithelium in vitro, ex vivo and in vivo. Gene Ther 2003; 10:1282–88 [DOI] [PubMed] [Google Scholar]
- 79.Adedoyin A, Bernardo JF, Swenson CE, Bolsack LE, Horwith G, DeWit S, Kelly E, Klasterksy J, Sculier JP, DeValeriola D, Anaissie E, Lopez-Berestein G, Llanos-Cuentas A, Boyle A, Branch RA. Pharmacokinetic profile of ABELCET (amphotericin B lipid complex injection): combined experience from phase I and phase II studies. Antimicrob Agents Chemother 1997; 41:2201–08 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Alton EWFW, Armstrong DK, Ashby D, Bayfield KJ, Bilton D, Bloomfield EV, Boyd AC, Brand J, Buchan R, Calcedo R, Carvelli P, Chan M, Cheng SH, Collie DDS, Cunningham S, Davidson HE, Davies G, Davies JC, Davies LA, Dewar MH, Doherty A, Donovan J, Dwyer NS, Elgmati HI, Featherstone RF, Gavino J, Gea-Sorli S, Geddes DM, Gibson JSR, Gill DR, Greening AP, Griesenbach U, Hansell DM, Harman K, Higgins TE, Hodges SL, Hyde SC, Hyndman L, Innes JA, Jacob J, Jones N, Keogh BF, Limberis MP, Lloyd-Evans P, Maclean AW, Manvell MC, McCormick D, McGovern M, McLachlan G, Meng C, Montero MA, Milligan H, Moyce LJ, Murray GD, Nicholson AG, Osadolor T, Parra-Leiton J, Porteous DJ, Pringle IA, Punch EK, Pytel KM, Quittner AL, Rivellini G, Saunders CJ, Scheule RK, Sheard S, Simmonds NJ, Smith K, Smith SN, Soussi N, Soussi S, Spearing EJ, Stevenson BJ, Sumner-Jones SG, Turkkila M, Ureta RP, Waller MD, Wasowicz MY, Wilson JM, Wolstenholme-Hogg P, Consortium U. 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; 3:684–91 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Stribling R, Brunette E, Liggitt D, Gaensler K, Debs R. Aerosol gene delivery in vivo. Proc Natl Acad Sci U S A 1992; 89:11277–81 LP- [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Alton EWFW, Middleton PG, Caplen NJ, Smith SN, Steel DM, Munkonge FM, Jeffery PK, Geddes DM, Hart SL, Williamson R, Fasold KI, Miller AD, Dickinson P, Stevenson BJ, McLachlan G, Dorin JR, Porteous DJ. Non–invasive liposome–mediated gene delivery can correct the ion transport defect in cystic fibrosis mutant mice. Nat Genet 1993; 5:135–42 [DOI] [PubMed] [Google Scholar]
- 83.Hyde SC, Gill DR, Higgins CF, Trezise AEO, MacVinish LJ, Cuthbert AW, Ratcliff R, Evans MJ, Colledge WH. Correction of the ion transport defect in cystic fibrosis transgenic mice by gene therapy. Nature 1993; 362:250–55 [DOI] [PubMed] [Google Scholar]
- 84.Bobo D, Robinson KJ, Islam J, Thurecht KJ, Corrie SR. Nanoparticle-Based medicines: a review of FDA-approved materials and clinical trials to date. Pharm Res 2016; 33:2373–87 [DOI] [PubMed] [Google Scholar]
- 85.Dong Y, Siegwart DJ, Anderson DG. Strategies, design, and chemistry in siRNA delivery systems. Adv Drug Deliv Rev 2019; 144:133–47 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Kauffman KJ, Webber MJ, Anderson DG. Materials for non-viral intracellular delivery of messenger RNA therapeutics. J Controll Rel 2016; 240:227–34 [DOI] [PubMed] [Google Scholar]
- 87.Peraro L, Kritzer JA. Emerging methods and design principles for cell-penetrant peptides. Angewand Chem Int 2018; 57:11868–81 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Walsh M, Tangney M, O'Neill MJ, Larkin JO, Soden DM, McKenna SL, Darcy R, O'Sullivan GC, O'Driscoll CM. Evaluation of cellular uptake and gene transfer efficiency of pegylated poly-l-lysine compacted DNA: implications for cancer gene therapy. Mol Pharm 2006; 3:644–53 [DOI] [PubMed] [Google Scholar]
- 89.Andaloussi SEL, Lehto T, Mäger I, Rosenthal-Aizman K, Oprea II, Simonson OE, Sork H, Ezzat K, Copolovici DM, Kurrikoff K, Viola JR, Zaghloul EM, Sillard R, Johansson HJ, Said Hassane F, Guterstam P, Suhorutšenko J, Moreno PMD, Oskolkov N, Hälldin J, Tedebark U, Metspalu A, Lebleu B, Lehtiö J, Smith CIE, Langel U. Design of a peptide-based vector, PepFect6, for efficient delivery of siRNA in cell culture and systemically in vivo. Nucl Acids Res 2011; 39:3972–87 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Behrendt R, White P, Offer J. Advances in fmoc solid-phase peptide synthesis. J Peptide Sci 2016; 22:4–27 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Regberg J, Vasconcelos L, Madani F, Langel Ü, Hällbrink M. pH-responsive PepFect cell-penetrating peptides. Int J Pharm 2016; 501:32–38 [DOI] [PubMed] [Google Scholar]
- 92.Shahbazi M-A, Bauleth-Ramos T, Santos HA. DNA hydrogel assemblies: bridging synthesis principles to biomedical applications. Adv Ther 2018; 1:1800042–42 [Google Scholar]
- 93.Fink TL, Klepcyk PJ, Oette SM, Gedeon CR, Hyatt SL, Kowalczyk TH, Moen RC, Cooper MJ. Plasmid size up to 20 kbp does not limit effective in vivo lung gene transfer using compacted DNA nanoparticles. Gene Ther 2006; 13:1048–51 [DOI] [PubMed] [Google Scholar]
- 94.Konstan MW, Davis PB, Wagener JS, Hilliard KA, Stern RC, Milgram LJH, Kowalczyk TH, Hyatt SL, Fink TL, Gedeon CR, Oette SM, Payne JM, Muhammad O, Ziady AG, Moen RC, Cooper MJ. Compacted DNA nanoparticles administered to the nasal mucosa of cystic fibrosis subjects are safe and demonstrate partial to complete cystic fibrosis transmembrane regulator reconstitution. Hum Gene Ther 2004; 15:1255–69 [DOI] [PubMed] [Google Scholar]
- 95.Ziady A-G, Gedeon CR, Miller T, Quan W, Payne JM, Hyatt SL, Fink TL, Muhammad O, Oette S, Kowalczyk T, Pasumarthy MK, Moen RC, Cooper MJ, Davis PB. Transfection of airway epithelium by stable PEGylated poly-l-lysine DNA nanoparticles in vivo. Mol Ther 2003; 8:936–47 [DOI] [PubMed] [Google Scholar]
- 96.Ziady A-G, Gedeon CR, Muhammad O, Stillwell V, Oette SM, Fink TL, Quan W, Kowalczyk TH, Hyatt SL, Payne J, Peischl A, Seng JE, Moen RC, Cooper MJ, Davis PB. Minimal toxicity of stabilized compacted DNA nanoparticles in the murine lung. Mol Ther 2003; 8:948–56 [DOI] [PubMed] [Google Scholar]
- 97.Young JL, Dean DA. Electroporation-mediated gene delivery. Adv Genet. 2015; 89:49–888 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Esmaeili N, Friebe M. Electrochemotherapy: a review of current status, alternative IGP approaches, and future perspectives. J Healthcare Eng 2019; 2019:2784516. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Geboers B, Scheffer HJ, Graybill PM, Ruarus AH, Nieuwenhuizen S, Puijk RS, van den Tol PM, Davalos RV, Rubinsky B, de Gruijl TD, Miklavčič D, Meijerink MR. High-voltage electrical pulses in oncology: irreversible electroporation, electrochemotherapy, gene electrotransfer, electrofusion, and electroimmunotherapy. Radiology 2020; 295:254–72 [DOI] [PubMed] [Google Scholar]
- 100.Krassowska W, Filev PD. Modeling electroporation in a single cell. Biophys J 2007; 92:404–17 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Saulis G, Saulė R. Size of the pores created by an electric pulse: microsecond vs millisecond pulses. Biochim Biophys Acta 2012; 1818:3032–39 [DOI] [PubMed] [Google Scholar]
- 102.Chandrasekaran AP, Song M, Kim K-S., Ramakrishna S. Different methods of delivering CRISPR/Cas9 into cells. Prog Mol Biol Transl Sci. 2018; 159:157–176 [DOI] [PubMed] [Google Scholar]
- 103.Chen S, Lee B, Lee AY-F, Modzelewski AJ, He L. Highly efficient mouse genome editing by CRISPR ribonucleoprotein electroporation of zygotes. J Biol Chem 2016; 291:14457–67 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Seki A, Rutz S. Optimized RNP transfection for highly efficient CRISPR/Cas9-mediated gene knockout in primary T cells. J Exp Med 2018; 215:985–97 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Dean DA. Electroporation of the vasculature and the lung. DNA Cell Biol 2003; 22:797–806 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Young JL, Barravecchia MS, Dean DA. Electroporation-mediated gene delivery to the lungs. In: Li S, Cutrera J, Heller R, Teissie J (eds) Electroporation protocols: preclinical and clinical gene medicine New York, NY: Springer, 2014, pp. 189–204 [DOI] [PubMed]
- 107.Campana LG, Edhemovic I, Soden D, Perrone AM, Scarpa M, Campanacci L, Cemazar M, Valpione S, Miklavčič D, Mocellin S, Sieni E, Sersa G. Electrochemotherapy – emerging applications technical advances, new indications, combined approaches, and multi-institutional collaboration. Eur J Surg Oncol 2019; 45:92–102 [DOI] [PubMed] [Google Scholar]
- 108.Gothelf A, Mir LM, Gehl J. Electrochemotherapy: results of cancer treatment using enhanced delivery of bleomycin by electroporation. Cancer Treat Rev 2003; 29:371–87 [DOI] [PubMed] [Google Scholar]
- 109.Ricke J, Jürgens JHW, Deschamps F, Tselikas L, Uhde K, Kosiek O, De Baere T. Irreversible electroporation (IRE) fails to demonstrate efficacy in a prospective multicenter phase II trial on lung malignancies: the ALICE trial. Cardiovasc Intervent Radiol 2015; 38:401–08 [DOI] [PubMed] [Google Scholar]
- 110.Vroomen LGPH, Petre EN, Cornelis FH, Solomon SB, Srimathveeravalli G. Irreversible electroporation and thermal ablation of tumors in the liver, lung, kidney and bone: what are the differences? Diagn Intervent Imag 2017; 98:609–17 [DOI] [PubMed] [Google Scholar]
- 111.Impellizeri JA, Ciliberto G, Aurisicchio L. Electro-gene-transfer as a new tool for cancer immunotherapy in animals. Veterinary Comparat Oncol 2014; 12:310–18 [DOI] [PubMed] [Google Scholar]
- 112.Pasquet L, Bellard E, Chabot S, Markelc B, Rols M-P, Teissie J, Golzio M. Pre-clinical investigation of the synergy effect of interleukin-12 gene-electro-transfer during partially irreversible electropermeabilization against melanoma. J Immunother Cancer 2019; 7:161–61 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Pasquet L, Chabot S, Bellard E, Markelc B, Rols M-P, Reynes J-P, Tiraby G, Couillaud F, Teissie J, Golzio M. Safe and efficient novel approach for non-invasive gene electrotransfer to skin. Sci Rep 2018; 8:16833. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Thalmensi J, Pliquet E, Liard C, Chamel G, Kreuz C, Bestetti T, Escande M, Kostrzak A, Pailhes-Jimenez A-S, Bourges E, Julithe M, Bourre L, Keravel O, Clayette P, Huet T, Wain-Hobson S, Langlade-Demoyen P. A DNA telomerase vaccine for canine cancer immunotherapy. Oncotarget 2019; 10:3361–72 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Bai H, Lester GMS, Petishnok LC, Dean DA. Cytoplasmic transport and nuclear import of plasmid DNA. Biosci Rep 2017; 37:BSR20160616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Dean DA, Machado-Aranda D, Blair-Parks K, Yeldandi AV, Young JL. Electroporation as a method for high-level nonviral gene transfer to the lung. Gene Ther 2003; 10:1608–15 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Machado-Aranda D, Adir Y, Young JL, Briva A, Budinger GRS, Yeldandi AV, Sznajder JI, Dean DA. Gene transfer of the Na+,K+-ATPase β1 subunit using electroporation increases lung liquid clearance. Am J Respir Crit Care Med 2005; 171:204–11 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Dean DA, Barravecchia M, Danziger B, Lin X. Use of electroporation for efficacious gene delivery to the lungs. ECS Trans 2011; 35:167–77 [Google Scholar]
- 119.Pringle IA, McLachlan G, Collie DDS, Sumner-Jones SG, Lawton AE, Tennant P, Baker A, Gordon C, Blundell R, Varathalingam A, Davies LA, Schmid RA, Cheng SH, Porteous DJ, Gill DR, Hyde SC. Electroporation enhances reporter gene expression following delivery of naked plasmid DNA to the lung. J Gene Med 2007; 9:369–80 [DOI] [PubMed] [Google Scholar]
- 120.Dolgachev V, Panicker S, Balijepalli S, McCandless LK, Yin Y, Swamy S, Suresh MV, Delano MJ, Hemmila MR, Raghavendran K, Machado-Aranda D. Electroporation-mediated delivery of FER gene enhances innate immune response and improves survival in a murine model of pneumonia. Gene Ther 2018; 25:359–75 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Dolgachev VA, Goldberg R, Suresh MV, Thomas B, Talarico N, Hemmila MR, Raghavendran K, Machado-Aranda D. Electroporation-mediated delivery of the FER gene in the resolution of trauma-related fatal pneumonia. Gene Ther 2016; 23:785–96 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Lin X, Barravecchia M, Kothari P, Young JL, Dean DA. β1-Na+,K+-ATPase gene therapy upregulates tight junctions to rescue lipopolysaccharide-induced acute lung injury. Gene Ther 2016; 23:489–89 [DOI] [PubMed] [Google Scholar]
- 123.Machado-Aranda D, Raghavendran K. Electroporation-mediated delivery of genes in rodent models of lung contusion. Meth Mol Biol 2014; 1121:205–21 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Machado-Aranda DA, Suresh MV, Yu B, Raghavendran K. Electroporation-mediated in vivo gene delivery of the Na+/K+-ATPase pump reduced lung injury in a mouse model of lung contusion. J Trauma Acute Care Surg 2012; 72:32–40 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Mutlu GM, Machado-Aranda D, Norton JE, Bellmeyer A, Urich D, Zhou R, Dean DA. Electroporation-mediated gene transfer of the Na+,K+ -ATPase rescues endotoxin-induced lung injury. Am J Respir Crit Care Med 2007; 176:582–90 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Gazdhar A, Fachinger P, van Leer C, Pierog J, Gugger M, Friis R, Schmid RA, Geiser T. Gene transfer of hepatocyte growth factor by electroporation reduces bleomycin-induced lung fibrosis. Am J Physiol 2007; 292:L529–L36 [DOI] [PubMed] [Google Scholar]
- 127.Gazdhar A, Temuri A, Knudsen L, Gugger M, Schmid RA, Ochs M, Geiser T. Targeted gene transfer of hepatocyte growth factor to alveolar type II epithelial cells reduces lung fibrosis in rats. Hum Gene Ther 2012; 24:105–16 [DOI] [PubMed] [Google Scholar]
- 128.Lin X, Barravecchia M, Matthew Kottmann R, Sime P, Dean DA. Caveolin-1 gene therapy inhibits inflammasome activation to protect from bleomycin-induced pulmonary fibrosis. Sci Rep 2019; 9:19643–43 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Barnett RC, Lin X, Barravecchia M, Norman RA, de Mesy Bentley KL, Fazal F, Young JL, Dean DA. Featured article: electroporation-mediated gene delivery of surfactant protein B (SP-B) restores expression and improves survival in mouse model of SP-B deficiency. Exp Biol Med 2017; 242:1345–54 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.DeGiulio JV, Kaufman CD, Dean DA. The SP-C promoter facilitates alveolar type II epithelial cell-specific plasmid nuclear import and gene expression. Gene Ther 2010; 17:541–49 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Gottfried L, Lin X, Barravecchia M, Dean DA. Identification of an alveolar type I epithelial cell-specific DNA nuclear import sequence for gene delivery. Gene Ther 2016; 23:734–42 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Deipolyi AR, Golberg A, Yarmush ML, Arellano RS, Oklu R. Irreversible electroporation: evolution of a laboratory technique in interventional oncology. Diagn Intervent Radiol 2014; 20:147–54 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Scheffer HJ, Nielsen K, de Jong MC, van Tilborg AAJM, Vieveen JM, Bouwman A, Meijer S, van Kuijk C, van den Tol P, Meijerink MR. Irreversible electroporation for nonthermal tumor ablation in the clinical setting: a systematic review of safety and efficacy. J Vasc Intervent Radiol 2014; 25:997–1011 [DOI] [PubMed] [Google Scholar]
- 134.Badding MA, Lapek JD, Friedman AE, Dean DA. Proteomic and functional analyses of protein-DNA complexes during gene transfer. Mol Ther 2013; 21:775–85 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Rolong A, Davalos RV, Rubinsky B. History of electroporation. In: Meijerink MR, Scheffer HJ, Narayanan G (eds) Irreversible electroporation in clinical practice. Cham: Springer International Publishing, 2018, pp. 13–37
- 136.Belur LR., McIvor RS. Delivery of transposon DNA to lungs of mice using polyethyleneimine-DNA complexes BT – gene delivery to mammalian cells. Vol. 1 In: Heiser WC (ed) Nonviral gene transfer techniques. Totowa: Humana Press, 2004. pp. 137–43 [DOI] [PubMed]
- 137.Muliawan Hs NKYK, et al. No title. Kobe J Med Sci 2015; 61:E47–E53 [PubMed] [Google Scholar]
- 138.Gill DR, Smyth SE, Goddard CA, Pringle IA, Higgins CF, Colledge WH, Hyde SC. Increased persistence of lung gene expression using plasmids containing the ubiquitin C or elongation factor1α promoter. Gene Ther 2001; 8:1539–46 [DOI] [PubMed] [Google Scholar]
- 139.Yew NS, Przybylska M, Ziegler RJ, Liu D, Cheng SH. High and sustained transgene expression in vivo from plasmid vectors containing a hybrid ubiquitin promoter. Mol Ther 2001; 4:75–82 [DOI] [PubMed] [Google Scholar]
- 140.Yew NS, Zhao H, Przybylska M, Wu IH, Tousignant JD, Scheule RK, Cheng SH. CpG-depleted plasmid DNA vectors with enhanced safety and Long-Term gene expression in vivo. Mol Ther 2002; 5:731–38 [DOI] [PubMed] [Google Scholar]
- 141.Chabot S, Bellard E, Reynes JP, Tiraby G, Teissie J, Golzio M. Electrotransfer of CpG free plasmids enhances gene expression in skin. Bioelectrochemistry 2019; 130:107343–43 [DOI] [PubMed] [Google Scholar]
- 142.Batrakova EV, Kim MS. Using exosomes, naturally-equipped nanocarriers, for drug delivery. J Controll Rel 2015; 219:396–405 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Costa Verdera H, Gitz-Francois JJ, Schiffelers RM, Vader P. Cellular uptake of extracellular vesicles is mediated by clathrin-independent endocytosis and macropinocytosis. J Controll Rel 2017; 266:100–08 [DOI] [PubMed] [Google Scholar]
- 144.El-Andaloussi S, Lee Y, Lakhal-Littleton S, Li J, Seow Y, Gardiner C, Alvarez-Erviti L, Sargent IL, Wood MJA. Exosome-mediated delivery of siRNA in vitro and in vivo. Nat Protoc 2012; 7:2112–26 [DOI] [PubMed] [Google Scholar]
- 145.Hou R, Li Y, Sui Z, Yuan H, Yang K, Liang Z, Zhang L, Zhang Y. Advances in exosome isolation methods and their applications in proteomic analysis of biological samples. Analyt Bioanalyt Chem 2019; 411:5351–61 [DOI] [PubMed] [Google Scholar]
- 146.Jeyaram A, Jay SM. Preservation and storage stability of extracellular vesicles for therapeutic applications. AAPS J 2017; 20:1–1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Lee M, Ban J-J, Im W, Kim M. Influence of storage condition on exosome recovery. Biotechnol Bioproc Eng 2016; 21:299–304 [Google Scholar]
- 148.Mathiyalagan P, Sahoo S. Exosomes-Based gene therapy for MicroRNA delivery. Meth Mol Biol 2017; 1521:139–52 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Pyne NJ, El Buri A, Adams DR, Pyne S. Sphingosine 1-phosphate and cancer. Adv Biol Regul 2017; 68:97–106 [DOI] [PubMed] [Google Scholar]
- 150.Chen W-X, Zhou J, Zhou S-S, Zhang Y-D, Ji T-y, Zhang X-L, Wang S-M, Du T, Ding D-G. Microvesicles derived from human Wharton’s jelly mesenchymal stem cells enhance autophagy and ameliorate acute lung injury via delivery of miR-100. Stem Cell Res Ther 2020; 11:113–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Park O, Choi ES, Yu G, Kim JY, Kang YY, Jung H, Mok H. Efficient delivery of tyrosinase related protein-2 (TRP2) peptides to lymph nodes using serum-derived exosomes. Macromol Biosci 2018; 18:1800301. [DOI] [PubMed] [Google Scholar]
- 152.Zhang Y, Böse T, Unger RE, Jansen JA, Kirkpatrick CJ, van den Beucken JJJP. Macrophage type modulates osteogenic differentiation of adipose tissue MSCs. Cell Tissue Res 2017; 369:273–286 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Public Safety Notification on Exosome Products from the FDA and published Dec 6, 2019 (https://www.fda.gov/vaccines-blood-biologics/safety-availability-biologics/public-safety-notification-exosome-products)
- 154.Usman WM, Pham TC, Kwok YY, Vu LT, Ma V, Peng B, Chan YS, Wei L, Chin SM, Azad A, He AB-L, Leung AYH, Yang M, Shyh-Chang N, Cho WC, Shi J, Le MTN. Efficient RNA drug delivery using red blood cell extracellular vesicles. Nature Communications 2018; 9:2359–59 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Exo-Fect Exosome Transfection Kit, System Biosciences. 2018. (https://systembio.com/shop/exo-fect-exosome-transfection-kit/).


