Next Article in Journal
Circulatory Serum Krebs von Den Lungen-6 and Surfactant Protein-D Concentrations Predict Interstitial Lung Disease Progression and Mortality
Next Article in Special Issue
Effects of Rho-Associated Kinase (Rock) Inhibitors (Alternative to Y-27632) on Primary Human Corneal Endothelial Cells
Previous Article in Journal
Bone Morphogenetic Protein-4 Impairs Retinal Endothelial Cell Barrier, a Potential Role in Diabetic Retinopathy
Previous Article in Special Issue
Pre-Clinical Evaluation of Efficacy and Safety of Human Limbus-Derived Stromal/Mesenchymal Stem Cells with and without Alginate Encapsulation for Future Clinical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Corneal Regeneration Using Gene Therapy Approaches

by
Subhradeep Sarkar
1,2,†,
Priyalakshmi Panikker
1,†,
Sharon D’Souza
3,
Rohit Shetty
3,
Rajiv R. Mohan
4,5,6 and
Arkasubhra Ghosh
1,*
1
GROW Research Laboratory, Narayana Nethralaya Foundation, Bangalore 560099, Karnataka, India
2
Manipal Academy of Higher Education, Manipal 576104, Karnataka, India
3
Department of Cornea and Refractive Surgery, Narayana Nethralaya, Bangalore 560010, Karnataka, India
4
Harry S. Truman Memorial Veterans’ Hospital, Columbia, MO 65201, USA
5
One-Health Vision Research Program, Departments of Veterinary Medicine and Surgery and Biomedical Sciences, College of Veterinary Medicine, University of Missouri, Columbia, MO 65211, USA
6
Mason Eye Institute, School of Medicine, University of Missouri, Columbia, MO 65211, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2023, 12(9), 1280; https://doi.org/10.3390/cells12091280
Submission received: 28 January 2023 / Revised: 13 April 2023 / Accepted: 23 April 2023 / Published: 28 April 2023

Abstract

:
One of the most remarkable advancements in medical treatments of corneal diseases in recent decades has been corneal transplantation. However, corneal transplants, including lamellar strategies, have their own set of challenges, such as graft rejection, delayed graft failure, shortage of donor corneas, repeated treatments, and post-surgical complications. Corneal defects and diseases are one of the leading causes of blindness globally; therefore, there is a need for gene-based interventions that may mitigate some of these challenges and help reduce the burden of blindness. Corneas being immune-advantaged, uniquely avascular, and transparent is ideal for gene therapy approaches. Well-established corneal surgical techniques as well as their ease of accessibility for examination and manipulation makes corneas suitable for in vivo and ex vivo gene therapy. In this review, we focus on the most recent advances in the area of corneal regeneration using gene therapy and on the strategies involved in the development of such therapies. We also discuss the challenges and potential of gene therapy for the treatment of corneal diseases. Additionally, we discuss the translational aspects of gene therapy, including different types of vectors, particularly focusing on recombinant AAV that may help advance targeted therapeutics for corneal defects and diseases.

1. Introduction

Corneal diseases are one of the leading causes of visual impairment globally [1]. A delicate physiological and functional balance is responsible for the transparency and clarity of the cornea. Therefore, any disease, acquired or genetic, that compromises this state of homeostasis can eventually lead to vision loss. Inherited, acquired, and iatrogenic corneal diseases, including corneal dystrophies, neovascularization, corneal scarring, haze, dry eyes, keratoconus, and corneal injuries, affect normal vision.
Approval of a gene therapy for retinal disease [2] has paved the way for corneal gene therapy; this method continues to develop and rapidly advance, and has a high potential for human application. Corneal gene therapy initially emerged in 1994, when corneal tissues were transduced successfully using replication-deficient adenovirus to treat acquired inflammatory disease [3]. Gene therapy for retinal diseases has made much greater progress when compared to corneal diseases. However, advancements made in the last two decades in the field of corneal gene therapy have brought this form of therapy much closer to clinical trials and application in human patients. Inherited corneal diseases, such as epithelial, stromal, and endothelial dystrophies [4,5], are natural candidates for corneal gene therapy. Complex conditions of multifactorial etiology, such as keratoconus, also have a genetic component associated to it, making it suitable for gene therapy [6].
The anatomic location of corneal epithelium makes it particularly attractive for non-invasive treatment by direct topical instillation of the gene delivery system [7]. Surgical, mechanical, chemical, and electric methods can also be used to administer gene therapy in the cornea. Furthermore, estimating the effectiveness and safety of the corneal treatment is easier due to the rapid and non-invasive visual observation using standard ophthalmologic methods [8]. Since the cornea is avascular and the eye has an altered immunologic profile in the intraocular compartments that usually dampens immune responses, the chances of systemic or even local immune reactions is lower.
The most striking advancement in the treatment of corneal diseases has been corneal transplantation and tissue engineering [9,10]. One of the most transplanted tissues worldwide is the cornea. In penetrating keratoplasty, the entire cornea is replaced, whereas in lamellar keratoplasty, only the damaged layers are replaced with requisite layers of donor tissue [10,11,12]. Unfortunately, in a subset of subjects, corneal transplantation has been associated with poor outcomes due to graft rejection and graft failure. Furthermore, the WHO has reported a shortage of corneal donors, due to which 10–15% of patients are reported to remain untreated [1,13]. Additionally, it has been reported that almost 53% of the world’s population lacks access to corneal transplantation, due to which there is only one cornea available for every seventy needed [14]. Due to the shortage of donor tissue and the fear of transmissible disease [14,15,16], there has been a continuous evolution of approaches in reviving corneal function and vision. An advantage of the cornea is its stability ex-vivo; it can be maintained in artificial physiological environment for weeks. This is especially advantageous for experimental purposes as well as for the administration of treatments prior to corneal transplantation, aiding in reduced chance of graft rejection [17,18].
To reduce global blindness, the development of a tissue-targeted gene-based intervention based on our understanding of molecular mechanisms that can be targeted by gene therapy is of paramount importance. Many of the diseases could be potentially treated using gene therapy by supplying a functional gene or changing the expression levels of specific genes in affected cellular layers. However, the majority of the studies in the field have been focused on the modulation of acquired corneal medical conditions. Regulation of the corneal microenvironment could be achieved using corneal gene therapy in different disorders by using induction or knock down of respective proteins. To improve the efficacy and safety of the treatment, local corneal gene delivery using a variety of vector and delivery modalities can be used to achieve low and continuous concentrations of the proteins [19]. Therefore, delivery of various immune modulators and growth factors to the cornea, in turn affecting local immune responses, inflammation, and proliferation, can be achieved without any systemic side effects. Corneal gene therapy has therefore been applied to the treatment of fibrotic disorders and corneal haze post-refractive procedures. Topical gene delivery to the cornea has the potential to effectuate local protein expression at concentrations unachievable by systemic administration of the transgenes or recombinant proteins [20]. Corneal gene therapy approaches can therefore enhance survival of corneal grafts and improve corneal diseases that currently require a corneal transplantation, thus preventing the need for an allograft [21].
In this review, we discuss genes associated with diverse corneal diseases that could serve as potential candidates for corneal gene therapy and the complexities linked with them. We also explore the strategies, challenges, and potential of gene therapy for the treatment of corneal diseases. Further, we discuss the multiple aspects of gene therapy that can help to improve personalized therapeutics in the field of corneal diseases.

2. Genes of Corneal Diseases

One of the first requirements for a successful gene-based treatment approach for the cornea is to identify the genes that need to be augmented, replaced, or edited, or identification of the pathways to be targeted by the protein product of the selected transgene.

2.1. Genes and Genetics of Inherited Diseases

Corneal dystrophies: Corneal dystrophies represent a heterogenous group of inherited corneal diseases that affect corneal development and cellular function. Corneal dystrophies are classified based on the layer of cornea that they affect. The human cornea essentially consists of the epithelium, the stroma, and the endothelium layer. Bowman’s layer, an acellular structure with densely compacted collagen fibrils, separates the epithelium and the anterior stroma. The Descemet’s membrane is the basement membrane of the endothelial layer that separates the posterior part of the stroma from the endothelium. Corneal dystrophies are classified into four main categories: epithelial and subepithelial dystrophies; Bowman’s layer dystrophies; stromal dystrophies; and Descemet’s membrane and endothelial dystrophies (Figure 1). Each of these dystrophies exhibit distinct clinical features, variable inheritance, age of onset, and course of progression [4,5] (Table 1). Corneal dystrophies are inherited as autosomal dominant, autosomal recessive, or X-linked.
TGFβI-associated dystrophies have been attributed to various mutants in TGFβI gene, with positions 124 and 555 being the most common. Deposition of TGFβ-induced (TGFβ) protein aggregates in the stroma or Bowman layer have been associated with these dystrophies [22]. Several uncommon mutations are scattered across five of the gene’s 17 exons and are associated with divergent clinical presentations of the dystrophy, including Reis–Bucklers corneal dystrophy, Thiel–Behnke corneal dystrophy, lattice corneal dystrophy type 1, granular corneal dystrophy type 1, and granular corneal dystrophy type 2. Epithelial recurrent erosion dystrophy (ERED) is caused by mutation in the COL17A1 gene [23]. Table 1 shows the current international classification of most corneal dystrophies. Furthermore, sequencing analyses showed that mutations in COL17A1 were causative of ERED [23]. Until the discovery of this mutation, Thiel–Behnke corneal dystrophy was thought to result only from mutations in the TGFBI gene. Heterozygous mutations in the KRT3 or KRT12 genes, encoding corneal specific Keratins 3 and 12, respectively [24], have been associated with Meesmann epithelial corneal dystrophy (MECD), a rare autosomal dominant inherited disease. Malfunction of K3 and K12 has been shown to cause mechanical fragility of the anterior corneal epithelium [25]. Various studies have identified at least 24 mutations for MECD, most of which are missense point mutation [24,26,27,28]. Early onset Fuchs endothelial corneal dystrophy (FECD) has been associated with mutations in the COL8A2, SLC4A11, ZEB1, and LOXHD1 genes. The majority of FECD cases are caused by a trinucleotide repeat expansion in the TCF4 gene [29], leading to altered mRNA processing due to sequestration of splicing factor proteins (MBNL1 and MBNL2) to the nuclear RNA foci.
Other inherited diseases: Other inherited disorders affecting the cornea include Aniridia and Mucopolysaccharidosis (MPS). Aniridia exhibits a dominant autosomal inheritance pattern with variable expression in several members of a family. A diverse set of mutations leading to haploinsufficiency of the PAX6 gene, which is expressed in various regions of the eye including the cornea [30], have been demonstrated to be associated this disorder. MPS is a group of inherited metabolic disorders caused by the absence or malfunctioning of lysosomal enzymes that are responsible for the degradation of glycosaminoglycans. Lysosomal accumulation of undegraded glycosaminoglycans in keratocytes causes corneal clouding. MPS demonstrate a high range of clinical manifestations and are classified based on the enzyme that is dysregulated. Although common to most types of MPS (I-IX), corneal manifestations are most common in MPS I, VI, and VII. MPS I is a monogenetic disease caused by loss of function mutations in both copies of the IDUA gene. MPS VII, or Sly syndrome, is caused by mutation in the enzyme β-glucuronidase [31]; MPS VI, or Maroteaux–Lamy syndrome, is caused by mutations leading to dysfunction of the enzyme aryl sulfatase B, causing corneal clouding [32].

2.2. Genes Associated with Acquired Corneal Conditions

2.2.1. Corneal Wound Healing

Disparate injuries can cause fibrotic scarring. Irrespective of the cause of the injury, transforming growth factor β (TGFβ) has been associated with aberrant corneal healing responses. Previous studies have demonstrated the interaction of the TGF-β superfamily proteins with the Smad family to activate downstream signaling [33]. Bone morphogenic protein is a part of TGF-β superfamily that appears to modulate keratocyte proliferation, differentiation, and apoptosis, thereby playing an essential role during corneal wound healing [34]. Studies have shown the important role of BMP7 in the development of the mammalian kidney and eye [35]. The complex wound healing signaling network involves the binding of BMP7 to type I and II receptors, allowing regulation of receptor-regulated Smads (Smad 1, 5, and 8) and inhibitory Smads (Smad 6 and 7) [36]. Improved healing and lessened scarring by overexpression of the inhibitory protein Smad7, thereby causing inhibition of TGFβ signaling pathway, have been observed in animal models [37]. Additionally, hepatocyte growth factor (HGF) and its receptor (c-Met receptor tyrosine kinase) have also been shown to play important roles in normal and healing corneal epithelium and keratocytes in the anterior stroma in vivo [38]. Upregulation of HGF in keratocytes has been observed upon injury to the corneal epithelium [39].

2.2.2. Corneal Neovascularization

Corneal neovascularization occurs in various corneal pathologies, including inflammatory diseases, congenital diseases, contact lens-related hypoxia, inflammatory diseases, autoimmune diseases, chemical burns, trauma, corneal graft rejection, and infectious keratitis [40], which may lead to significant visual impairment or blindness [41]. Various angiogenic factors, such as VEGF, basic fibroblast growth factor (bFGF), matrix metalloproteinase (MMP), platelet-derived growth factors (PDGFs), and interleukin-1 (IL-1), mediate corneal neovascularization. In the retina and the cornea, VEGF-A is the predominant VEGF member driving pathologic neovascularization and is therefore a potential target of several drugs. The VEGF-A binds to two members of receptor tyrosine kinase family, VEGF receptors VEGFR1 and VEGFR2, also known as Flt-1 and KDR, respectively. Vascular endothelial cells (VECs) significantly express both of these receptors, and their expression increases in the presence of inflammation. It has been found that their activation promotes vascular leakage, VEC liberation, and VEC proteolysis [42,43]. Targeting these factors and inhibiting their expression or increasing expression of anti-angiogenic factors may help inhibit angiogenesis. Bevacizumab and ranibizumab are currently used in clinical settings to inhibit VEGF-A signaling pathways [44]. When mixed with non-liposomal lipid and delivered by means of subconjunctival injection, the brain-specific angiogenesis inhibitor 1 (BAI1-ECR) gene showed an effective reduction of corneal neovascularization [45]. Several of these targets that mediate neovascularization, including Flt23k, Flt-1, PEDF, VEGFR Flt-1, MMP-9, and vasohibin-1 [46,47,48,49,50,51], could serve as a potential target candidates for corneal gene therapy.

2.2.3. Corneal Graft Survival

Immunological rejection has always been one of the primary causes for corneal graft rejection [52]. Several different groups have shown significant prolongation of corneal allograft survival using various different transgenes and vectors in a variety of animal models [53]. Several of these transgenes lower immune response [54,55,56,57,58] and reduce angiogenesis, thereby preventing likelihood of donor graft tissue rejection due to inflammation, corneal scarring, and edema [59]. The number of transgenes that have shown success in modulating corneal graft rejection indicates the complex nature of the process and the multiple pathways involved. Prevention of activation of T cells by gene transfer of Cytotoxic T-Lymphocyte Antigen 4 protein (CTLA4-Ig) has shown to effectively prolong graft survival [56,60]. Furthermore, knocking down neuropilin-2 through RNA interference (RNAi) [61] has been shown to reduce the amount of free vascular endothelial growth factor-A (VEGF-A) [62], leading to a decrease in activated T-cell influx and improvement in graft survival.

2.2.4. Multifactorial and Polygenic Diseases

A variety of conditions are associated with reduced central corneal thickness (CCT) which include keratoconus, keratoglobus, brittle cornea syndrome, Ehlers–Danlos syndrome, osteogenesis imperfecta, and myopia [63]. CCT reduction is also associated with various genetic determinants in the context of these diseases, although various systemic and environmental factors are also involved in disease. The most common among these conditions is keratoconus (KC), a complex multifactorial disease associated with a wide variety of etiological factors such as genetic predisposition, environmental insults, oxidative stress, and mechanical injuries or eye rubbing. Alterations in several biochemical mediators have been associated to KC pathogenesis, including lysyl oxidase (LOX) [64,65], MMP2 [66], MMP9 [65,67], and various collagens, including I, III, IV, V, VI, and VII [68]. Familial genetic studies and genome-wide studies of the KC families have identified genomic heterogeneity in the genomic loci among these families [69,70,71,72,73]. Mutations in DOCK9, FLG, TGFβI, SOD1, ZEB1, and VSX1 genes were also found to be responsible for KC prognosis in some selected populations around the world [74,75,76,77,78,79].

3. Clinical Presentation of Various Corneal Diseases

Corneal disease is a major public health problem primarily attributed to infections, trauma, corneal scars, corneal dystrophies, and degenerative and multifactorial conditions [80]. Corneal dystrophies are a group of inherited diseases that can involve various layers of the cornea and are usually bilateral and progressive, resulting in progressive loss of the transparency of the cornea and visual deterioration [81]. They affect around 0.09% of the global population, with a majority of cases being of endothelial origin [82]. These conditions usually present at the clinic with a progressive decrease in vision and can also come with additional problems, such as severe photophobia and irritation due to recurrent corneal erosions (RCE) seen in the more superficial corneal involvement. Slit lamp examination aids the clinical diagnosis by the classical features observed. Episodes of RCE can have a deleterious effect on the quality of life of these patients and can also result in visual deterioration due to subepithelial or stromal scarring.
The deposition of various substances in the cornea also results in early visual impairment requiring surgical intervention. Certain dystrophies, such as the congenital stromal dystrophy and congenital hereditary endothelial dystrophy (CHED), result in amblyopia in these patients [83]; in Fuch’s endothelial dystrophy, there is intrastromal fluid collection, causing corneal scarring in advanced disease [84,85].
One of the most common dystrophic corneal conditions is Keratoconus, an ectatic disease characterized by the thinning of the corneal stroma; the disease is usually bilateral and can be progressive in nature. It has a typical onset in the first to second decade of age and can progress until the third to fourth decade. As the disease progresses, there is a significant deterioration in vision due to the irregular astigmatism and ectasia [86]. In advanced stages, there can be varying degrees of corneal scarring and even rupture of the Descemet’s membrane, resulting in corneal hydrops [87].
Corneal scars can develop after infections, trauma, and certain surgeries, including refractive procedures such as photorefractive keratectomy (PRK), or collagen crosslinking for KC [88,89]. This results in suboptimal visual outcomes.

4. Clinical Treatment Options Currently Available

The treatment of corneal dystrophies depends on the degree of involvement of the cornea and stage of the disease. Surgical treatment options include phototherapeutic keratectomy (PTK) in more superficial involvement and lamellar or full thickness corneal transplantation in more advanced disease.
Various modalities of treatment are available for other corneal conditions, such as KC, depending on the stage of the disease. Collagen cross linking is the primary method of management for progressive KC and can be combined with adjunctive procedures such as topography guided ablation and intracorneal ring segments. In advanced KC with scarring, treatment options include specialty contact lenses such as the scleral lenses, which are effective but may not be tolerated by all patients, or a lamellar or full thickness corneal transplant depending on the depth and extent of scarring [90]. With a prevalence of 1.38 per 1000 [91], many patients are at risk of becoming permanently blind. In addition, since the patients are typically affected during their productive age of 15–40 years, the socio-economic burden of KC is very high [92]. Furthermore, patients with KC experience a significant impact on their quality of life. A higher probability of having a psychiatric disorder has been associated with more severe cases of KC.
Visual rehabilitation for corneal scars includes the use of scleral contact lenses, photorefractive or phototherapeutic keratectomy using an excimer laser, or a keratoplasty, depending on the depth and position of the scar [88,93]. When the scar involves the entire corneal thickness, it requires a penetrating keratoplasty, while scars or opacities limited to a section of the cornea can be managed by a lamellar keratoplasty of the anterior or posterior cornea as required [11].

5. Current Challenges in the Treatment of Corneal Diseases

The available surgical options are effective in treating these conditions but do pose certain clinical and logistical challenges. Phototherapeutic keratectomy (PTK) is a useful and effective technique using excimer laser ablation to treat superficial corneal opacities secondary to infections, trauma, or dystrophy [10]. However, there can be recurrence of the primary pathology post-procedure, especially in the case of corneal dystrophies. Other possible complications post-procedure include induced refractive errors, infections, and corneal scarring [94].
One of the most important methods for treating various corneal conditions is corneal transplantation. Lamellar techniques of corneal transplantation have markedly improved outcomes for corneal grafts [95]. However, despite the rising popularity of these transplantation techniques, there are still several challenges faced with corneal transplants. A major problem faced in eye banking and corneal transplantation is the availability of viable corneal tissue for transplantation [80]. There is large disparity between the tissue requirement and the supply of corneal tissues, with a more than 4-fold difference between available tissue for transplantation and actual required numbers [14,96]. There is also the rejection of potential donor tissues during the screening process due to transmissible diseases such HIV and Hepatitis B and C [15,16]. Additionally, it is estimated that more than 50% of the world’s population does not have access to the surgical expertise required for corneal transplantation [1].
Another challenge faced in corneal transplantation is the possibility of graft rejections and failure. Although the early graft survival of full thickness corneal transplant is more than 70% 1 year post surgery, this survival rate deteriorates to less than 50% at 5 years, even in primary grafts [97,98]. This is further reduced in repeat grafts and high-risk grafts. Lamellar techniques of corneal transplantation, such as DALK, DSEK, and DMEK, have less risk of rejection and better surgical and visual outcomes in low-risk grafts [99,100]. In high-risk grafts, there is loss of the cornea’s immune privilege due to various preexisting conditions; treatment may involve long term immunomodulation with attendant complications and vary despite it [99,101]. In developing nations, there is also the problem related to cost of the procedure, long term follow-ups, and the availability of surgical and technical expertise [102,103].

6. Gene Therapy for Corneal Diseases

Gene therapy is a technique in which replacing or inactivating (knocking out) the mutated gene occurs as a result of targeted therapeutic delivery of correct nucleic acid directly into the patient’s cells. Gene therapy to treat any inherited corneal dystrophies works via three approaches: (a) Gene inactivation or silencing of the mutated gene that manifests a toxic effect on the cells; (b) Gene correction or replacement of the mutated gene with a normal copy of a healthy gene; and (c) Addition of a healthy copy of the gene that will ectopically express the therapeutic protein to rescue the disease phenotype. Gene therapy has a growing potential in the field of corneal dystrophies due to three major characteristics: accessibility in terms of injections and surgical interventions; its partial immune-privileged properties that limit immune responses towards the antigenicity of the transgene and the viral vectors; and the presence of a tight blood-retinal barrier that can help to prevent unintentional spreading or contamination of the neighboring tissues as well as to the general circulation. In addition, direct corneal delivery, either topically or via injection, can be achieved only in the affected eye in the case of unilateral conditions. The advancement in imaging technologies, such as pentacam and adaptive optics, further facilitates quantitative and qualitative evaluation of corneal changes after gene therapy. However, the wide heterogeneity in disease-causing genes in hereditary corneal disease requires knowledge and identification of the specific causative gene in each patient in order to consider gene therapy as an intervention. However, many of the pathways affected in the tissues as part of their pathology can be common. Therefore, there is a need to design differential strategies according to the causative genes, mutation, and inheritance pattern for gene therapy of various hereditary corneal dystrophies. The success of gene therapy for the treatment of any disease depends on the efficiency by which the therapeutic transgene is delivered to the target cell type. There are currently two main approaches in terms of vectors for delivering genetic material: viral and non-viral vectors.

6.1. Viral Vectors

Significant progress in understanding the molecular mechanism of diverse corneal diseases has led to the development of gene therapies in animal models. Gene therapy in animal models has demonstrated restoration of the biomechanical stability of the cornea by regulating the key proteins that are deficient in various corneal dystrophies, thus establishing a scientific basis for application in human subjects. Most studies on gene therapy of diverse corneal dystrophies have used viral vectors, including adeno-associated virus (AAV), lentivirus, and adenovirus, due to their unique ability to transduce a wide tropism of living cells with minimal resistance. The mechanisms by which the viral vectors deliver the therapeutic gene and translate the therapeutic protein into the host cells is discussed in this section.

6.1.1. Adeno-Associated Virus (AAV)

AAV is a small (25-nm), non-pathogenic, nonenveloped virus that packages a linear single-stranded DNA genome of ~4.7 kb containing two genes that produce 4 Rep and 3 Cap proteins for replication and capsid proteins, respectively (Figure 2). The Rep gene encodes for four different proteins: rep40, rep52, rep68, and rep78. Rep68 and rep78 play an essential role in viral genome integration, replication, and transcriptional regulation of AAV gene expression, whereas rep 40 and rep 52 proteins are involved in viral genome encapsulation. The cap gene encodes three viral capsid proteins, VP1 (90 kDa), VP2 (72 kDa), and VP3 (60 kDa) (Figure 2), which are arranged in a 1:1:10 ratio and form an icosahedral symmetry. AAV belongs to the family Parvoviridae and is placed in the genus Dependovirus [104], as AAV requires a helper virus, such as adenovirus, to establish a productive infection cycle. However, recombinant AAV vectors used for gene therapy do not carry any of its native Rep/Cap genes or any helper genes, underlining its safety profile for therapeutic applications. Additionally, these vectors can transduce and express in both mitotic and post-mitotic cells.
The ability of the various rAAV serotypes to transduce a wide variety of ocular structures is due to its ability to bind primary cell surface receptors such as heparin sulphate proteoglycan (HSPGs) [105]. HSPGs are expressed in most of the cell types and uses integrin αvβ5 or fibroblast growth factor receptor (FGFR) as coreceptors for its internalization and endocytosis [106,107]. Once endocytosed, the viral particles are released from the endosome at a low pH [108,109]. The released SS-DNA of AAV is converted to ds-DNA by either annealing to another complementary strand of AAV or via the host cell DNA replication machinery. Subsequently, the viral genome remains as an episome and, with the help of host cell machinery, undergoes transcription and translation to produce the deficient therapeutic protein in the target cells over extended durations (Figure 2). This non-integrative nature of AAV vectors is another key safety feature wherein the risk of insertional mutagenesis in the host is low. A limitation of AAV vectors are their low cargo carrying capacity (Table 2).
The ability of the various rAAV serotypes to transduce ocular structures at the anterior segment of the eye has been extensively documented using vectors encoding marker proteins; therefore, it has become evident that a combination of serotypes, route of administration, and the choice of regulatory elements such as promotors allows the selective tropism of desired cell types. Recent studies have also shown that the corneal stroma is an ideal target for AAV-mediated gene therapy where the quiescent stromal keratocytes can receive the vectors and express the therapeutic proteins over long durations. In the context of corneal gene therapy, the first reported gene therapy was mediated by the AAV2 serotype, which demonstrated successful transgene delivery into a rabbit cornea in vivo [110]. The natural occurrence of the AAV2 serotypes may cause its pre-exposure in humans. This exposure leads to the production of neutralizing antibodies against the serotype. Thus, when a therapeutic gene packaged into a AAV2 serotype is delivered to the patient there is a chance of eliciting a humoral immune response against it which in turn may diminish the efficacy and safety of viruses carrying the therapeutic gene. The AAV5 serotype has been found to be the most divergent serotype [111], with a several-fold more efficient transduction efficiency than AAV2 [112]. In vivo gene therapy for corneal scarring using topical administration of the AAV5 serotype in rabbit models has shown a significant decrease in corneal haze and fibrosis, without any reports of any immunogenic or toxic immune response against it [113,114] (Table 3). In the domain of AAV vectors, AAV8 [115] and AAV9 [116] were considered the most efficient in corneal keratocyte transduction. Therefore, a chimeric AAV capsid was generated where an AAV8 capsid scaffold was engrafted with the AAV9 galactose binding domain. On evaluating the potency of infection of the chimeric AAV8/9 capsid following an intrastromal injection of the vectors into the human donor eyeballs, efficient widespread transduction of the transgene was observed. The AAV8/9 chimeric capsids were found to have greater and more efficient transduction efficiency when compared to either parent serotype at similar vector doses [117]. Therefore, a direct gene augmentation strategy using AAV vectors offers a viable option to cure diseases for which no treatment options exist.

6.1.2. Lentivirus (LV)

Lentiviruses are single-stranded enveloped RNA viruses and are members of the retroviridae family. A mature LV is 100 nm in diameter and has a cylindrical core structure. LV transduction takes place through association with specific cell surface receptors. After binding to the receptor, the viral membrane fuses with the host cell membrane [126] and injects the nucleoprotein complex into the cell. After its entry, the viral RNA is reverse-transcribed into ds-DNA by the reverse transcriptase enzyme associated with the nucleoprotein complex. The ds-DNA enters the nucleus through the nuclear pore complex and is integrated into the host genome with the help of the viral integrase [127] (Figure 3). The major merit of using LVs is its larger transgene carrying capacity, broad tropism, and its ability to transduce both dividing and non-dividing cells (Table 2). Numerous ex vivo gene therapy experiments using LVs for corneal graft rejection and corneal fibrosis (Table 4) in various animal models have shown successfully enhanced transgene expression in the corneal epithelium, endothelium, and keratocytes and thus rescuing the disease phenotype. The cons of using LVs as a vector for corneal gene therapy are that it possesses high immune cell infiltration post-infection, random integration potential [128,129], and a consequent risk of insertional mutagenesis/teratoma formation. Due to these crucial shortcomings, the use of LVs still requires immense evaluation and identification of preferential integration sites in the host genome before being applicable to humans for therapies.

6.1.3. Adenovirus (Ad)

Adenoviruses are double-stranded enveloped DNA viruses and are members of the Adenoviridae family. More than 50 known human adenoviral serotypes are present in nature, of which serotypes 2 and 5 are the most widely used in gene therapy. Ad vectors can transduce both dividing and non-dividing cells. It can harbor a therapeutic gene up to 30Kb in size and deliver it to the target tissue. Their inability to integrate into the host genome reduces the chances of insertional mutagenesis. In adenoviral vector preparations, high titers of pure viruses can be easily obtained from a single viral prep, thus a small volume of vector injection into the cornea is enough for a high level of transduction. These characteristic features of Ads make them an attractive candidate for gene therapy studies (Table 2). The Ad vectors deliver the therapeutic gene in the host cell through receptor-mediated endocytosis [135], by binding to the cell surface coxsackie adenovirus receptor, αvβ3 integrin, and clathrin-coated pits. Once internalized, the viral genome containing the therapeutic gene is released into the cytoplasm which further enters the nucleus through the nuclear pore complex and remains distinct from the host genome and is expressed as episomes to produce the therapeutic protein (Figure 3). Successful ex vivo and in vivo gene therapy of the cornea was first investigated by adenoviral vectors in different animal models (Table 5). However, the shortcomings were the short-lived expression of the delivered transgene, primarily due to the rapid histone association and silencing of the Ad genome. This necessitates repeated vector injection, which showed more toxic responses to the cells than the original exposure [136]. The Ad capsid structures (the pentons and heptons) are potent targets for rapid immune recognition causing Ad mediated gene therapy to typically induce significant host immune response [137]. Also, most humans are pre-exposed to wild-type adenovirus and therefore the immune responses are also mounted against the viruses, thereby giving a pre-existing immunity to the cells [17]. This ultimately dampens the efficacy of the vectors intended for the therapy (Table 2).

6.2. Non-Viral Vectors

The delivery of the therapeutic gene into the cornea can also be achieved using a non-viral origin vector mediated gene therapy. This includes the use of liposomes, compacted nanoparticles, electroporation, particle gun bombardment, and many other methods. The advantage of non-viral vectors is low immunogenicity, ease of manipulating their chemical properties to suit DNA delivery, the possibility of large-scale production, and transferring large vectors without any immune reactions. However, non-viral vectors may present obstacles such as inefficient cellular membrane transport, intracellular/extracellular degradation, and lack of long-term gene expression [146].

6.2.1. Electroporation

Electroporation is a technique that uses short and intense electric pulses to create pores or reversibly permeabilize the cell membrane in order to deliver the naked circular DNA carrying the therapeutic gene into the target cells (Figure 4). This physical method of non-viral mediated gene therapy has been widely used for successfully delivering plasmid DNA into the corneal endothelium and stromal keratocytes [147,148]. The optimal field strength of 100–200 V/cm was found to be suitable for the transfer of naked plasmid DNA without any corneal damage or inflammation. Various in vivo experiments on different animal models for stromal keratitis and corneal endothelium wound healing experiments showed a 1000-fold increase in gene uptake in the cornea in comparison to injection of DNA alone, leading to inhibition of the disease phenotype [20,149]. One of the main drawbacks when using electroporation is that the permeabilization as a result of the electric pulse becomes irreversible due to the heat generated during the process [150].

6.2.2. Nanoparticles

Nanoparticles are ultrafine particles that range between 1 and 100 nm in size and are widely used in nanomedicine due to their (a) small size, (b) ability to deliver the transgene into the intracellular compartment of the cells, (c) high surface area to volume ratio, (d) capability to deliver a larger payload, and (e) negligible toxic damage to the cell membrane. The use of nanoparticles is highly suitable for the delivery of transgene for the treatment of eye-related diseases as they can permeabilize across various ocular barriers including the cornea, sclera, conjunctiva, and, in some cases, the blood-retinal barriers. Nanoparticles can harbor multiple cargo types, including DNA, peptides, antibodies, molecular sensors, and drugs, in the desired cellular layers of the eye. The feasibility of using nanoparticles in various in vivo experiments to rescue the disease phenotype in corneal scarring has been demonstrated by several groups (Table 6). Nanoparticles are mainly classified as metallic, polymeric, and hybrid nanoparticles. Polymeric nanoparticles that are usually prepared from albumin, chitosan, and polyethyleneimine (PEI) have been found to be more efficient in delivering the transgene into rodent corneas in vivo without any significant side effects [151,152,153]. In vivo delivery of transgene using hybrid nanoparticles in rabbit cornea to treat corneal fibrosis has demonstrated an efficient cargo of transgene in the target cells with significant inhibition of fibrosis and no visible toxic effects [36].
The direct transfer of the therapeutic gene to the corneal cells in vivo has been achieved using cationic lipids [154]. The positively charged cationic lipids bind to negatively charged DNA molecules to form a lipid-DNA complex that has a high affinity for the cell membrane. The lipid-DNA complex is then endocytosed into an endocytic vesicle followed by its trafficking and release from the endosomal compartment. The nuclear uptake of the DNA takes place through the nuclear pore complex, which then forms an episome to express the therapeutic protein using the host cell machinery [155] (Figure 4). The effectiveness and safety of liposome-mediated therapeutic gene delivery in various ex vivo and in vitro studies has been demonstrated in the last few years (Table 6).
Table 6. Corneal in vivo and ex vivo gene therapy using non-viral vectors.
Table 6. Corneal in vivo and ex vivo gene therapy using non-viral vectors.
Non-Viral Mediated Corneal Gene Therapy
VectorDiseaseGenePromoterDosageModelSpeciesMode of AdministrationOutcomeReference
ElectroporationStromal KeratitisIL-10CMV, UbC1 µL Plasmid solution (5 µg Plasmid DNA in 10 mM Tris, pH 8.0; 1 mM EDTA; and 140 mM NaCl)In VivoFemale Balb/c mice weighing 16 to 24 gStromal injection of plasmid DNA, followed by gold-plated Genetrode electrodes were placed on the cornea on either side of the area injected. An ECM 830 square wave electroporator was used to deliver eight pulses of 10-msec duration at a field strength of 200 V/cm.Gene expression driven by the CMV promoter remained high for three days, which started to reduce by 2-fold each day thereafter. Replacing the promoter with UbC surprisingly showed similar half-life gene expression. Also, an adverse effect was observed when using DNA nuclear-targeting sequences in vectors.[148]
CRISPR/dCas9Corneal Endothelial wound healingSOX2Not mentioned0.1 nmolIn Vivo6–8 weeks old Sprague–Dawley rats; a 12-h light/dark cycle at 25 °C for 7 days before initiating experiments.Anterior chamber injection of plasmid DNA, followed by 7-mm Tweezertrodes were placed on each cornea, with the positive electrode on the plasmid-injected eye. The parameters were set at 140 V, 100 milliseconds length, 950 milliseconds interval, five pulses, and 100 V/cm2SOX2 activation promoted the reduction of central corneal thickness and corneal opacity in comparison to the control group. Additionally, an increase in Cell viability, proliferation rate, and the number of cells in the S-phase was observed after SOX2 overexpression.[147]
CRISPR-Cas9Granular corneal dystrophy (GCD 2)TGFBINot mentionedCRISPR-Cas9 constructs (2.5 µg per well) and ssODN (1 µg per well)In VitroHuman, GCD 2 patient-derived corneal keratocytesTransfectionEffective gene correction efficiency of R124H mutation associated with GCD2 disorder was observed without any off-target effects. In heterozygous cells, the correction efficiency was found to be 20.6% and in homozygous 41.3% respectively.[156]
LipofectionMeesmann’s Epithelial Corneal Dystrophy (MECD)KRT12Not mentioned200 ng plasmid (1 well of 12 well plate) In VitroCorneal limbal epithelial cell derived from limbal biopsy of MECD patientsLipofectamine 2000Potent and specific knockdown of K12-Leu132Pro at both the mRNA and protein levels. An allele-specific knockdown of 63% of the endogenous mutant allele was observed.[157]
Lipofectionlattice corneal dystrophy type I (LCDI)TGFBI-Arg124CysCMV200 ng plasmid (1 well of 12 well plate)Ex VivoCorneal limbal epithelial cell derived from limbal biopsy of LCD1 patientsLipofectamine 2000The siRNA specific to TGFBI-Arg124Cys, efficiently suppressed the mutant allele[158]
PEI gold nanoparticleCorneal FibrosisBMP7CAG37.5 µL of 150 mM PEI2-GNPs with 10 µg of plasmid DNAIn VivoFemale New Zealand White rabbit weighing 2.0 to 3.0 kgTopicalSignificant inhibition of fibrosis post-PRK was observed. [36]
PEI nanoparticleCorneal FibrosisDecorinCAG150 mM linear 22 kDa PEI+ 2 µg of plasmidIn VitroHorseTopical22 KDa PEI nanoparticle effectively inhibited TGFb-mediated fibrosis[159]

6.3. Gene Therapy Strategies for Precise and Targeted Therapeutics

6.3.1. Gene Augmentation

Gene augmentation therapy is a straightforward approach in which the transfer of the functional therapeutic gene into the affected cells takes place to restore the expression of an inadequately functioning gene [160,161] (Figure 5). This approach is mainly applicable to recessive genetic diseases. For the augmentation strategy to be effective, the augmented transgene must synthesize physiological and adequate levels of the normal protein and the tissue status and disease stage at which treatment is attempted should not be terminal. To minimize off-target transgene expression that may have a pathogenic effect within the cells, it is important to focus on the appropriate regulation of the transgene expression for augmentation-based approaches. To address appropriate transgene regulation, cell or tissue-specific promoter sequences can be designed to modulate the expression of the therapeutic transgene. For example, in response to inflammation, NFκB responsive promoter sequences have been used to drive the transgene expression in immune organs [162]. A commonly used approach in gene augmentation therapy for secretory protein is to target the delivery of the therapeutic gene to a distant tissue/cell and not the affected cell type or tissue; this will help in the production of the same deficient therapeutic secretory protein to rescue the disease phenotype. This approach can be adopted when (i) transduction of affected cells with the therapeutic gene may not be able to produce a physiologically relevant amount of secreted transgene protein; (ii) the affected site of gene delivery has been rendered sensitive to any further damage that can be caused during the gene delivery procedure; and (iii) the site of gene delivery is highly exposed to the host immune system which can lead to a transgene-directed immune response. Such an approach in corneal gene therapy has not yet been listed in the literature.
Gene augmentation therapy for corneal diseases has been conducted in various animal models in vivo to rescue the disease phenotype. There are numerous studies that have shown that gene augmentation using viral and non-viral vectors can help to significantly improve the disease pathology in murine, ovine, and canine models.

6.3.2. Gene Editing

In recent times, gene editing has become one of the most utilized approaches in the field of gene therapy. This approach is able to target both autosomal dominant and recessively inherited corneal dystrophies that lead to loss of function due to mutation of the target gene. The strategy aims to restore/correct the normal functioning of the mutated gene by adding, removing, or altering the genome at the site of the mutation. One of the most specific and advanced genome editing technologies to date that has been used for gene correction in the anterior segment of the eye is Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR) and CRISPR-associated (Cas) proteins, termed the CRISPR-Cas system [163]. Meganucleases, Zinc finger nucleases (ZFNs) [164], and transcription activator-like effector nucleases (TALENs) are also used in this approach. CRISPR-Cas-based genome editing has been found to be more favorable than the other mentioned nucleases due to its relatively simple design for targeting the mutated gene [165]. On the other hand, significant modifications in the DNA-binding protein domains are required to target with ZFNs and TALENs.
The mechanism of CRISPR-Cas 9 involves the binding of Cas9 nuclease at the genomic locus that is complementary to the guide RNA; this guide RNA creates a double-stranded break, which is repaired via two mechanisms: (i) error-prone non-homologous end joining (NHEJ) or (ii) homologous directed repair (HDR) (Figure 5). NHEJ can create insertions or deletions in the mutated gene, resulting in the formation of a premature stop codon. This can be used to knock out the expression of the truncated protein. HDR can help incorporate specific alterations into the mutated gene provided by the repair template. All studied Cas9 enzymes require a protospacer adjacent motif (PAM) next to the target site for target DNA recognition. Some recent studies have also shown the ability of certain divergent CRISPR-Cas9 enzymes that can recognize and cleave single-stranded DNA (ssDNA) using an RNA-guided, PAM-independent recognition mechanism [166]. Intravenous delivery of the CRISPR-Cas9 system cannot be used in the eye due to the blood-ocular barrier [167]. AAV as a vector has been widely used for the delivery of CRISPR-Cas9 construct into the target tissue/cell. However, the biggest shortcoming of using AAV vectors is their carrying capacity. They are often too small to accommodate the full genome editing system. To address this issue, a dual-AAV vector design has been implemented in which the Cas9 nuclease and the single-guide RNA (sgRNA) cassettes have been packaged in two separate vectors and delivered to the target tissue, which showed highly efficient genome editing in the hepatocyte cells [168]. The CRISPR-Cas 9 system has been used to treat corneal dystrophy in animal models. In one study, selective disruption of the mutated copy of the KRT12 gene in a humanized MECD mouse model was achieved by targeting the protospacer adjacent motif (PAM) sequence that was generated by the missense mutation in the KRT12 gene [169].

6.3.3. Gene Silencing

Gene augmentation therapy is likely to be ineffective in rescuing the disease phenotype in autosomal dominant diseases. For augmentation therapy to work effectively, the suppression of the mutated gene expression must be addressed first. Silencing the mutated gene can be mediated via delivering a small double-stranded non-coding RNA construct that is designed to act via RNA interference (RNAi) [170]. RNAi promotes post-transcriptional gene silencing by enzymatic degradation of the complementary RNA species. This silencing is mediated by a large multi-component RNA/protein complex called the RNA-induced silencing complex (RISC) [171] (Figure 5). siRNAs (small interfering RNA) have been tested in various corneal dystrophies, including MECD, wound healing, and neovascularization [157,172,173].
In the context of mRNA-based gene therapy, the most successful approach is the use of antisense oligonucleotides (AONs). AONs are short single-stranded DNA or RNA that interacts with the complementary mRNA to block the translation by altering the pre-mRNA splicing [174]. The first AON approved by the FDA was for the treatment of cytomegalovirus retinitis: fomivirsen (Vitravene). In the cornea, the first phase clinical trial was conducted to explore the efficacy of a topically administered AON (Aganirsen) that targets the insulin substrate-1 receptor to block corneal neovascularization in keratitis patients. The AON administration was found effective and thus reduced the need for corneal transplantation [175]. GS-101 AON was also found to be a potent anti-angiogenic compound that helped in the significant regression of corneal neovascularization [176].
Despite gaining success, the strategy of gene silencing fails to completely silence the target gene expression. For the RNAi pathway, the major drawbacks, such as off-the-mark effects and extended toxicity of the small RNA molecules, must be considered and studied carefully for its successful application in clinical settings.

6.3.4. Dual Vectors

Limitations of using rAAV-based gene therapy correctional strategies include its limited packaging size for genes larger than 4.7 kb such as PD-L1, which is widely used in corneal graft rejection studies. Several strategies have been designed to take advantage of the head-tail concatamerization of the AAV genome (Figure 6). One strategy is to use the overlapping sequence approach. In this method, efficient reconstitution and transgene expression rely on the use of two separate (dual) AAV vectors, each one carrying half of a large gene. Upon coinfection of the target cell from both vectors, the two halves will be reconstituted due to the canonical ability of AAV genomes to concatamerize via intermolecular recombination. In 2007, Ghosh et al. [177] developed a hybrid dual vector strategy to expand the packaging capacity in the rAAV system, the trans-splicing vectors, and showed effective, whole-body transduction using a mouse model of Duchenne Muscular Dystrophy. This trans-splicing strategy employs the use of a splice donor at the 3′ end of one-half of the target gene in one vector and a splice acceptor at the 5′ end of the other half of the transgene. This allows for efficient reconstitution of the mRNA transcripts by the head-to-tail concatamerization process. The third dual AAV approach (hybrid) is a combination of the two previous approaches; it is based on the addition of a highly recombinogenic exogenous sequence (recombinogenic region) to the trans-splicing vectors in order to increase their recombination efficiency. In 2011, Ghosh et al. [178] reported the use of minimized bridging sequences from the highly recombinogenic alkaline phosphatase (~227–247 bp) to circumvent the available space for packaging the transgene. In this approach, the authors demonstrated the complete reconstitution of the B-Galactosidase gene (LacZ) when packaged separately and co-infected to check expression. Expression of the reconstituted LacZ was proven to be better than the original hybrid strategy developed in suitable cell cultures and animal models.

6.4. Current Scenario of Corneal Gene Therapy

In the last two decades, research in the field of ocular gene therapy has advanced from a conceptual validation to a clinical reality with the approval of the first vision restoring gene therapy treatment for LCA (Leber congenital amaurosis) [2]. However, advances made in research related to the cornea have been considerably less when compared to the retina [179]. Progress in the field of corneal gene therapy has been steadily progressing despite remaining at the preclinical level. Several studies have identified various therapeutic genes and used several animal models to demonstrate the use of gene therapy in treating corneal defects [180,181].
Most of the preclinical work has focused mainly on treating and preventing acquired conditions such as corneal haze, corneal wound healing, herpes simplex keratitis (HSK), corneal neovascularization, and corneal graft rejection (Table 3, Table 4, Table 5 and Table 6). Studies have shown that moderate corneal conditions, such as thick haze and inflammation produced by keratectomy surgery with excimer laser effectively, can be treated by gene therapies in rabbits in vivo [114,182]. Inhibitor of differentiation 3 (Id3) gene, a transcriptional repressor, is known to inhibit differentiation of corneal keratocyte to myofibroblasts. Recently, a study showed the therapeutic effects of AAV5.Id3 gene therapy on corneal pathology and ocular health in rabbit eyes with corneal scarring/fibrosis induced by alkali trauma [125]. Further, dual gene therapy with BMP7 and HGF effectively treated both fibrosis and neovascularization in a severe corneal opacity model produced by chemical injury [182]. Neovascularization in the host cornea is one of the primary reasons for corneal transplant rejection and therefore several studies have focused on reducing or preventing corneal neovascularization. TGFβ and VEGF growth factors play an important role in promoting fibrosis and neovascularization in the cornea in vivo. Decorin, a small leucine-rich proteoglycan, is a potent inhibitor of both these growth factors. AAV5 mediated decorin gene therapy has shown to effectively ameliorate corneal neovascularization and fibrosis in rabbit eyes in vivo [159,181,183]. Additionally, AAV5-Smad7 gene therapy has shown to inhibit corneal fibrosis post-PRK in rabbit stroma in vivo [113]. One study used a recombinant adeno-associated viral (rAAV) vector carrying an endostatin gene as an anti-angiogenic strategy to show successful inhibition of neovascularization when administrated by subconjunctival injection. Here, a minimal immune response with stable transgene expression was reported for 8 months [184]. Another study showed an effective reduction of corneal neovascularization when the brain-specific angiogenesis inhibitor 1 (BAI1-ECR) gene, mixed with non-liposomal lipid, was delivered by means of subconjunctival injection in an in vivo rabbit model [45].
The time between donor cornea collection and recipient corneal transplant would be a perfect time to apply a pre-treatment to the cornea to reduce the chance of corneal graft rejection. Recently, rabbit and human donor corneal buttons were incubated ex vivo with an AAV vector-mediated human leukocyte antigen G (HLA-G). The treated group did not exhibit edema and neovascularization for more than 2.5 months upon allotransplantation. Furthermore, xenotransplantation (human donor to rabbit recipient) showed a delayed rejection time from 18 days to 29 days [121]. Delivery of IL-10 and IL-12 with adenoviral vectors to ovine corneas exhibited higher rates of graft survival [55,144,145,185]. Increased endothelial cell survival was observed with lentiviral vector-mediated gene delivery of baculoviral p35 or mammalian Bcl-xL at various stages of storage [186]. With continued investigations, gene therapy has the potential to improve survival of corneal grafts as well as overcoming the need for corneal transplantation.
HSK is considered a major determinant of corneal graft rejection after transplantation. Most approaches have focused on preventing the development of herpetic lesions and diseases by downregulating the viral gene expression and affecting viral replication. More recently, LAT-targeting ribozymes were delivered using AAV to block viral reactivation in the eyes of rabbits with latent HSV-1 infection. Viral activation was blocked in more than 60% of the infected eyes [187]. Another study employed a different approach to target latent HSV-1 by using rare-cutting endonucleases such as meganucleases, which can be engineered for increased specificity. They delivered anti-HSV 1 meganucleases using rAAV to human cornea ex vivo before transplantation to reduce the chance of reinfection and graft rejection [188].
Gene therapy for treating patients with MPS has been under investigation for the last three decades. While phase I/II gene therapy trials are ongoing for some types of MPS (MPS I, II, IIA, IIIB, and VI) [189], the influence of these therapies on corneal opacity with MPS remains unclear. There are several gene therapy-based strategies targeting the ocular manifestation of MPS, including both systemic and local approaches, but these investigations are limited to animal models. Potential prevention and reversal of corneal blindness was shown using AAV8G9-IDUA (an AAV8 capsid scaffold with AAV9 putative galactose binding domain) gene therapy [117,120,190]. Additionally, reduced corneal clouding was seen with adenoviral-mediated expression of β-glucuronidase (GUSB) in the stromal region in MPS VII mouse models [191]. Similar results were obtained following adenovirus-mediated expression of human GUSB in canine MPS VII models [192]. Intravenous administration of AAV2/8 ARSB (arylsulfatase B) showed a favorable outcome in the liver of an MPS VI feline model, but the cornea was not evaluated [193].
Despite a wealth of knowledge on the methods for the delivery of transgene to the human cornea and ever-increasing information on the genes associated with various corneal dystrophies with substantial therapeutic potential, gene therapy for corneal dystrophies is yet to be explored.

7. Tailored Therapeutics Using Gene Therapy

Various approaches can be considered in order to increase precise and targeted expression of a therapeutic gene. The transgene cassette can be optimized to increase AAV transduction efficiency, vector tropism can be improved using capsid engineering, and an appropriate mode of delivery can be used to minimize off-target expression (Figure 7). Further, the capsid and transgene can be genetically modified to avoid the host immune response and the large-scale production of AAV can also be optimized.

7.1. Promoter Selection for Targeted Therapeutics

Small and tissue-specific promoters are important tools for preclinical research, for clinical delivery of gene therapies to avoid unwanted transgene expression at off-target areas, and to ensure cell type-specific gene expression. An effective promoter is important to pilot high and clinically relevant levels of therapeutic gene expression. The tissue specific promoter should allow a convenient vector dose for treatment without immune response or cellular toxicity resulting from high virus dosage. In preclinical research, tissue-specific promoters have been used to rescue many animal models of ocular disease. Studies have shown increased efficacy and safety by limiting unwanted off-target effects using tissue specific promoters. Tissue specific promoters may be considered from either the target gene itself or an unrelated gene with the appropriate expression pattern for the therapy.
Over the past few decades most gene therapy studies have used broad expression promoters such as cytomegalovirus (CMV), chicken beta-actin promoter (CAG), and human ubiquitin C promoter (UbiC). Currently, a small number of promoters specific to the cornea are being used, such as keratin K12 promoter and keratocan promoter [194,195]. Previously, promoters of several keratin genes have been used to drive tissue-specific expression of transgenes in animals. There are almost 30 different keratin proteins identified, of which keratin K12 and K3 are expressed in differentiated and stratified corneal epithelium. A study had showed high functional and tissue specific activity with the use of three 5′ truncated fragments of the keratin K12 promoter in human corneal epithelial cell lines [195]. The stroma of the cornea consists of many cell types, most of the cells being keratocytes [196]. Keratocan is a cornea stroma-specific keratan sulfate proteoglycan (KSPG) expressed in adult vertebrate cornea [195]. Li et al. identified the 3.2kb 5′ flanking promoter region of the keratocan gene and showed that it was able to drive β-galactosidase reporter specifically in the stromal layer of the cornea in adult transgenic mice [197]. Similarly, another study showed that keratocan promoter was capable of driving EGFP expression tissue-specifically in the corneal keratocyte [194].
More strongly regulated control of the protein expression level is required to avoid off target toxicity. This demand paved the way for the development of a variety of synthetic promoters in an effort to target expression in particular corneal cell types of interest as well as to provide physiologic expression of the exogenous transgenes. Recently, a group bioinformatically designed human DNA MiniPromoter (Ple253) for tissue-specific expression of transgene in the corneal stroma. Expression using Ple253 (PITX3) was robust and enriched in corneal stroma after neonatal intravenous delivery at P0. Furthermore, there was no expression in the brain, spinal cord, and heart but moderate expression in the liver and the pancreas. Expression at moderate levels in both the ganglion cell layer (GCL) and inner nuclear layer (INL) was observed after intravenous delivery at both P0 and P4 [198]. Additionally, there are several genes that have restricted expression in different regions of the cornea and could serve as a potential candidate promoter to drive cell-specific expression of a gene of interest. These include keratin-3 from the epithelium, decorin from the stroma [181], and ovary-specific protein from the endothelium [199]. Other potential candidate promoters include Aquaporin 5 (AQP5), which expressed in salivary and lacrimal glands and in corneal epithelium, and AQP1, expressed in corneal endothelium [200].
Another group demonstrated the possibility of designing cell-specific promoters in silico for in vivo applications. However, this was for retinal pigment epithelium (RPE) specific expression of the transgene [201]. Nevertheless, these findings demonstrate that with the increase in the quality and volume of ‘omics data and with the progressive development of TFRE database and informatic tools, it is now possible to successfully design synthetic promoters that will facilitate the advancement of more targeted and precise gene expression.

7.2. Capsid Engineering

One of the promising strategies to improve the efficacy of the AAV vector system for clinical application is capsid engineering. Various capsid engineering approaches include rational design and in-silico bioinformatic approaches [202]. The differences in protein sequence and structure between various wildtype capsids can lead to differences in transduction efficiency and the cell surface receptors utilized for entry. Further, it can also affect the relative biodistribution as well as the affinity for antibodies. Improving transduction efficiency will, in turn, help reduce the vector dose, lower the risk of host immune responses, and reduce the cost of manufacturing. This concept led to the idea of modifying capsid proteins to improve challenges associated with gene delivery. However, most of the studies to date have focused on modifying capsid to improve transduction efficiency in the retinal layers with little work conducted to improve transduction in the cornea.
Alteration of capsid to improve function involves rational design using the current information of the capsids including its crystal structures [203,204,205], cell surface receptors [205,206], immune system activation [207], infectious pathways [108], and antibody binding epitopes [208,209,210]. Experiments have shown an increase in AAV transduction efficiency with the presence of inhibitors of tyrosine kinases [211,212]. Additionally, it was shown that phosphorylation of tyrosine residue led to capsid degradation [213]. Zhong et al. showed that Y444F and Y730F mutations decreased phosphorylation and subsequent ubiquitination of the capsids, leading to significant improvement in transduction in vitro and in vivo. Another study showed improved transduction by mutating three surface tyrosine residues in murine hepatocytes in vivo [214]. Capsid modifications have also been extensively used to improve transduction efficiency in the retina. A study showed efficient transduction of mouse photoreceptors when four surface tyrosine residues and one surface threonine residue were mutated (Y272F, Y444F, Y500F, Y730F, T491V) [215].
Using emerging tools and technologies, continuous efforts are being made to advance capsid engineering. Furthermore, newer strategies to develop novel capsids are being used including next generation sequencing, AI-based machine learning, and bioinformatic capsid prediction tools that will dramatically accelerate progress towards achieving improved gene delivery and targeted therapeutics.

8. Challenges and Safety Aspect of Gene Therapy for Corneal Diseases

Despite significant advancement over the last two decades, more research is needed to address the challenges associated with corneal gene therapy. The risk of host immune responses towards the vector remains the largest challenge for AAV-based gene therapies. The capsid protein can elicit an immune response and lead to generation of neutralizing antibodies that could prevent the vector from infecting the patient cells and reduce the effectiveness of the treatment; however, given the immune advantaged location of the cornea, this is less of a risk. Furthermore, highly tissue specific expression of the transgene that will increase efficacy and safety by limiting unwanted off-target effect is required. However, across studies, the expression of transgenes is usually lower in corneal tissues compared to muscle, liver, etc., if normalized for dosage, therefore necessitating careful dose escalation studies. High dosage can add to the existing layer of complexity, where to overcome the barrier of delivering the right amount of AAV to targeted cells, higher dose therapies could cause safety concerns in the long term, particularly in the case of acquired diseases. Therefore, special consideration must be given to deliver vectors with “tunable” or “on/off” control expression cassettes. The last and major challenge associated with corneal gene therapy is the high cost of its research, development, and manufacturing. Further evaluation of the appropriate routes of administration, capsid choice, and vector genome designs are still required to advance corneal gene therapy from the preclinical to clinical setting.

9. Conclusions

Gene therapy approaches stand on the front line of advanced biomedical research and have matured considerably in the last decade as a new branch of regenerative medicine. Aside from corneal transplants, very few approaches are available to treat corneal diseases. Thus, gene therapy can be considered as a potential approach in treating various corneal conditions as it is able to correct the underlying pathological mechanism of a disease process with prolonged benefits. Identification of several genes involved in various inherited and acquired corneal conditions have further paved the way for the use of gene therapy approaches as a treatment modality for corneal conditions, including corneal neovascularization, corneal fibrosis, corneal graft rejection, and corneal dystrophies. However, further evaluation of the appropriate routes of administration, capsid choice, and vector genome designs are still needed to advance corneal gene therapy from the preclinical to clinical setting as a more precise and targeted therapeutic approach.

Author Contributions

Conceptualization, R.S. and A.G.; methodology, S.S., P.P. and A.G.; resources, R.S., S.D. and A.G.; data curation, S.S., P.P. and S.D.; writing—original draft preparation, S.S. and P.P.; writing—review and editing, R.R.M. and A.G.; supervision, A.G.; funding acquisition, R.S. and A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported India-Alliance DBT/Wellcome Trust grant to A.G (IA/TSG/20/1/6000029) and by the Narayana Nethralaya Foundation, Bangalore, India. The funders had no role in the analysis or writing of the manuscript.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

No patient data was used in the manuscript.

Data Availability Statement

Data will be available on request.

Acknowledgments

We gratefully acknowledge the contribution of Swaminathan Sethu at GROW research laboratory, Narayana Nethralaya Foundation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Whitcher, J.P.; Srinivasan, M.; Upadhyay, M.P. Corneal blindness: A global perspective. Bull. World Health Organ. 2001, 79, 214–221. [Google Scholar] [PubMed]
  2. Prado, D.A.; Acosta-Acero, M.; Maldonado, R.S. Gene therapy beyond luxturna: A new horizon of the treatment for inherited retinal disease. Curr. Opin. Ophthalmol. 2020, 31, 147–154. [Google Scholar] [CrossRef] [PubMed]
  3. Mashhour, B.; Couton, D.; Perricaudet, M.; Briand, P. In vivo adenovirus-mediated gene transfer into ocular tissues. Gene Ther. 1994, 1, 122–126. [Google Scholar] [PubMed]
  4. Constantin, C. Corneal dystrophies: Pathophysiological, genetic, clinical, and therapeutic considerations. Rom. J. Ophthalmol. 2021, 65, 104–108. [Google Scholar] [CrossRef] [PubMed]
  5. Lisch, W.; Weiss, J.S. Early and late clinical landmarks of corneal dystrophies. Exp. Eye Res. 2020, 198, 108139. [Google Scholar] [CrossRef]
  6. Loukovitis, E.; Sfakianakis, K.; Syrmakesi, P.; Tsotridou, E.; Orfanidou, M.; Bakaloudi, D.R.; Stoila, M.; Kozei, A.; Koronis, S.; Zachariadis, Z.; et al. Genetic Aspects of Keratoconus: A Literature Review Exploring Potential Genetic Contributions and Possible Genetic Relationships with Comorbidities. Ophthalmol. Ther. 2018, 7, 263–292. [Google Scholar] [CrossRef]
  7. Sonoda, S.; Tachibana, K.; Uchino, E.; Okubo, A.; Yamamoto, M.; Sakoda, K.; Hisatomi, T.; Sonoda, K.H.; Negishi, Y.; Izumi, Y.; et al. Gene transfer to corneal epithelium and keratocytes mediated by ultrasound with microbubbles. Investig. Ophthalmol. Vis. Sci. 2006, 47, 558–564. [Google Scholar] [CrossRef]
  8. Stechschulte, S.U.; Joussen, A.M.; von Recum, H.A.; Poulaki, V.; Moromizato, Y.; Yuan, J.; D’Amato, R.J.; Kuo, C.; Adamis, A.P. Rapid ocular angiogenic control via naked DNA delivery to cornea. Investig. Ophthalmol. Vis. Sci. 2001, 42, 1975–1979. [Google Scholar]
  9. Huang, Y.-X.; Li, Q.-H. An active artificial cornea with the function of inducing new corneal tissue generation in vivo—A new approach to corneal tissue engineering. Biomed. Mater. 2007, 2, S121. [Google Scholar] [CrossRef]
  10. Jhanji, V.; Mehta, J.S.; Sharma, N.; Sharma, B.; Vajpayee, R.B. Targeted corneal transplantation. Curr. Opin. Ophthalmol. 2012, 23, 324–329. [Google Scholar] [CrossRef]
  11. Chamberlain, W.D. Femtosecond laser-assisted deep anterior lamellar keratoplasty. Curr. Opin. Ophthalmol. 2019, 30, 256–263. [Google Scholar] [CrossRef] [PubMed]
  12. Luengo-Gimeno, F.; Tan, D.T.; Mehta, J.S. Evolution of deep anterior lamellar keratoplasty (DALK). Ocul. Surf. 2011, 9, 98–110. [Google Scholar] [CrossRef] [PubMed]
  13. Klausner, E.A.; Peer, D.; Chapman, R.L.; Multack, R.F.; Andurkar, S.V. Corneal gene therapy. J. Control. Release 2007, 124, 107–133. [Google Scholar] [CrossRef] [PubMed]
  14. Gain, P.; Jullienne, R.; He, Z.; Aldossary, M.; Acquart, S.; Cognasse, F.; Thuret, G. Global Survey of Corneal Transplantation and Eye Banking. JAMA Ophthalmol. 2016, 134, 167–173. [Google Scholar] [CrossRef] [PubMed]
  15. Fuest, M.; Yam, G.H.; Peh, G.S.; Mehta, J.S. Advances in corneal cell therapy. Regen. Med. 2016, 11, 601–615. [Google Scholar] [CrossRef]
  16. Mobaraki, M.; Abbasi, R.; Omidian Vandchali, S.; Ghaffari, M.; Moztarzadeh, F.; Mozafari, M. Corneal Repair and Regeneration: Current Concepts and Future Directions. Front. Bioeng. Biotechnol. 2019, 7, 135. [Google Scholar] [CrossRef]
  17. Mohan, R.R.; Rodier, J.T.; Sharma, A. Corneal gene therapy: Basic science and translational perspective. Ocul. Surf. 2013, 11, 150–164. [Google Scholar] [CrossRef]
  18. Torrecilla, J.; Del Pozo-Rodriguez, A.; Vicente-Pascual, M.; Solinis, M.A.; Rodriguez-Gascon, A. Targeting corneal inflammation by gene therapy: Emerging strategies for keratitis. Exp. Eye Res. 2018, 176, 130–140. [Google Scholar] [CrossRef]
  19. Sakamoto, T.; Oshima, Y.; Nakagawa, K.; Ishibashi, T.; Inomata, H.; Sueishi, K. Target gene transfer of tissue plasminogen activator to cornea by electric pulse inhibits intracameral fibrin formation and corneal cloudiness. Hum. Gene Ther. 1999, 10, 2551–2557. [Google Scholar] [CrossRef]
  20. Oshima, Y.; Sakamoto, T.; Yamanaka, I.; Nishi, T.; Ishibashi, T.; Inomata, H. Targeted gene transfer to corneal endothelium in vivo by electric pulse. Gene Ther. 1998, 5, 1347–1354. [Google Scholar] [CrossRef]
  21. Williams, K.A.; Jessup, C.F.; Coster, D.J. Gene therapy approaches to prolonging corneal allograft survival. Expert Opin. Biol. Ther. 2004, 4, 1059–1071. [Google Scholar] [CrossRef] [PubMed]
  22. Lakshminarayanan, R.; Chaurasia, S.S.; Anandalakshmi, V.; Chai, S.M.; Murugan, E.; Vithana, E.N.; Beuerman, R.W.; Mehta, J.S. Clinical and genetic aspects of the TGFBI-associated corneal dystrophies. Ocul. Surf. 2014, 12, 234–251. [Google Scholar] [CrossRef] [PubMed]
  23. Jonsson, F.; Bystrom, B.; Davidson, A.E.; Backman, L.J.; Kellgren, T.G.; Tuft, S.J.; Koskela, T.; Ryden, P.; Sandgren, O.; Danielson, P.; et al. Mutations in collagen, type XVII, alpha 1 (COL17A1) cause epithelial recurrent erosion dystrophy (ERED). Hum. Mutat. 2015, 36, 463–473. [Google Scholar] [CrossRef] [PubMed]
  24. Irvine, A.D.; Corden, L.D.; Swensson, O.; Swensson, B.; Moore, J.E.; Frazer, D.G.; Smith, F.J.; Knowlton, R.G.; Christophers, E.; Rochels, R.; et al. Mutations in cornea-specific keratin K3 or K12 genes cause Meesmann’s corneal dystrophy. Nat. Genet. 1997, 16, 184–187. [Google Scholar] [CrossRef] [PubMed]
  25. Omary, M.B.; Coulombe, P.A.; McLean, W.H. Intermediate filament proteins and their associated diseases. N. Engl. J. Med. 2004, 351, 2087–2100. [Google Scholar] [CrossRef]
  26. Hassan, H.; Thaung, C.; Ebenezer, N.D.; Larkin, G.; Hardcastle, A.J.; Tuft, S.J. Severe Meesmann’s epithelial corneal dystrophy phenotype due to a missense mutation in the helix-initiation motif of keratin 12. Eye 2013, 27, 367–373. [Google Scholar] [CrossRef]
  27. Nishida, K.; Honma, Y.; Dota, A.; Kawasaki, S.; Adachi, W.; Nakamura, T.; Quantock, A.J.; Hosotani, H.; Yamamoto, S.; Okada, M.; et al. Isolation and chromosomal localization of a cornea-specific human keratin 12 gene and detection of four mutations in Meesmann corneal epithelial dystrophy. Am. J. Hum. Genet. 1997, 61, 1268–1275. [Google Scholar] [CrossRef]
  28. Szaflik, J.P.; Oldak, M.; Maksym, R.B.; Kaminska, A.; Pollak, A.; Udziela, M.; Ploski, R.; Szaflik, J. Genetics of Meesmann corneal dystrophy: A novel mutation in the keratin 3 gene in an asymptomatic family suggests genotype-phenotype correlation. Mol. Vis. 2008, 14, 1713–1718. [Google Scholar]
  29. Wieben, E.D.; Aleff, R.A.; Tang, X.; Butz, M.L.; Kalari, K.R.; Highsmith, E.W.; Jen, J.; Vasmatzis, G.; Patel, S.V.; Maguire, L.J.; et al. Trinucleotide Repeat Expansion in the Transcription Factor 4 (TCF4) Gene Leads to Widespread mRNA Splicing Changes in Fuchs’ Endothelial Corneal Dystrophy. Investig. Ophthalmol. Vis. Sci. 2017, 58, 343–352. [Google Scholar] [CrossRef]
  30. Jordan, T.; Hanson, I.; Zaletayev, D.; Hodgson, S.; Prosser, J.; Seawright, A.; Hastie, N.; van Heyningen, V. The human PAX6 gene is mutated in two patients with aniridia. Nat. Genet. 1992, 1, 328–332. [Google Scholar] [CrossRef]
  31. Tomatsu, S.; Montano, A.M.; Dung, V.C.; Grubb, J.H.; Sly, W.S. Mutations and polymorphisms in GUSB gene in mucopolysaccharidosis VII (Sly Syndrome). Hum. Mutat. 2009, 30, 511–519. [Google Scholar] [CrossRef] [PubMed]
  32. Garrido, E.; Cormand, B.; Hopwood, J.J.; Chabas, A.; Grinberg, D.; Vilageliu, L. Maroteaux-Lamy syndrome: Functional characterization of pathogenic mutations and polymorphisms in the arylsulfatase B gene. Mol. Genet. Metab. 2008, 94, 305–312. [Google Scholar] [CrossRef]
  33. Massague, J.; Wotton, D. Transcriptional control by the TGF-beta/Smad signaling system. EMBO J. 2000, 19, 1745–1754. [Google Scholar] [CrossRef] [PubMed]
  34. Kim, W.J.; Mohan, R.R.; Mohan, R.R.; Wilson, S.E. Effect of PDGF, IL-1alpha, and BMP2/4 on corneal fibroblast chemotaxis: Expression of the platelet-derived growth factor system in the cornea. Investig. Ophthalmol. Vis. Sci. 1999, 40, 1364–1372. [Google Scholar]
  35. Dudley, A.T.; Lyons, K.M.; Robertson, E.J. A requirement for bone morphogenetic protein-7 during development of the mammalian kidney and eye. Genes Dev. 1995, 9, 2795–2807. [Google Scholar] [CrossRef]
  36. Tandon, A.; Sharma, A.; Rodier, J.T.; Klibanov, A.M.; Rieger, F.G.; Mohan, R.R. BMP7 gene transfer via gold nanoparticles into stroma inhibits corneal fibrosis in vivo. PLoS ONE 2013, 8, e66434. [Google Scholar] [CrossRef] [PubMed]
  37. Saika, S.; Ikeda, K.; Yamanaka, O.; Miyamoto, T.; Ohnishi, Y.; Sato, M.; Muragaki, Y.; Ooshima, A.; Nakajima, Y.; Kao, W.W.; et al. Expression of Smad7 in mouse eyes accelerates healing of corneal tissue after exposure to alkali. Am. J. Pathol. 2005, 166, 1405–1418. [Google Scholar] [CrossRef]
  38. Wilson, S.E.; Chen, L.; Mohan, R.R.; Liang, Q.; Liu, J. Expression of HGF, KGF, EGF and receptor messenger RNAs following corneal epithelial wounding. Exp. Eye Res. 1999, 68, 377–397. [Google Scholar] [CrossRef]
  39. Li, J.F.; Duan, H.F.; Wu, C.T.; Zhang, D.J.; Deng, Y.; Yin, H.L.; Han, B.; Gong, H.C.; Wang, H.W.; Wang, Y.L. HGF accelerates wound healing by promoting the dedifferentiation of epidermal cells through beta1-integrin/ILK pathway. Biomed. Res. Int. 2013, 2013, 470418. [Google Scholar] [CrossRef]
  40. Chang, J.H.; Gabison, E.E.; Kato, T.; Azar, D.T. Corneal neovascularization. Curr. Opin. Ophthalmol. 2001, 12, 242–249. [Google Scholar] [CrossRef]
  41. Kvanta, A.; Sarman, S.; Fagerholm, P.; Seregard, S.; Steen, B. Expression of matrix metalloproteinase-2 (MMP-2) and vascular endothelial growth factor (VEGF) in inflammation-associated corneal neovascularization. Exp. Eye Res. 2000, 70, 419–428. [Google Scholar] [CrossRef] [PubMed]
  42. Klettner, A.; Roider, J. Treating age-related macular degeneration—Interaction of VEGF-antagonists with their target. Mini Rev. Med. Chem. 2009, 9, 1127–1135. [Google Scholar] [CrossRef] [PubMed]
  43. Penn, J.S.; Madan, A.; Caldwell, R.B.; Bartoli, M.; Caldwell, R.W.; Hartnett, M.E. Vascular endothelial growth factor in eye disease. Prog. Retin. Eye Res. 2008, 27, 331–371. [Google Scholar] [CrossRef] [PubMed]
  44. Stevenson, W.; Cheng, S.F.; Dastjerdi, M.H.; Ferrari, G.; Dana, R. Corneal neovascularization and the utility of topical VEGF inhibition: Ranibizumab (Lucentis) vs bevacizumab (Avastin). Ocul. Surf. 2012, 10, 67–83. [Google Scholar] [CrossRef] [PubMed]
  45. Yoon, K.C.; Ahn, K.Y.; Lee, J.H.; Chun, B.J.; Park, S.W.; Seo, M.S.; Park, Y.G.; Kim, K.K. Lipid-mediated delivery of brain-specific angiogenesis inhibitor 1 gene reduces corneal neovascularization in an in vivo rabbit model. Gene Ther. 2005, 12, 617–624. [Google Scholar] [CrossRef]
  46. Cho, Y.K.; Uehara, H.; Young, J.R.; Tyagi, P.; Kompella, U.B.; Zhang, X.; Luo, L.; Singh, N.; Archer, B.; Ambati, B.K. Flt23k nanoparticles offer additive benefit in graft survival and anti-angiogenic effects when combined with triamcinolone. Investig. Ophthalmol. Vis. Sci. 2012, 53, 2328–2336. [Google Scholar] [CrossRef]
  47. Lai, C.M.; Shen, W.Y.; Brankov, M.; Lai, Y.K.; Barnett, N.L.; Lee, S.Y.; Yeo, I.Y.; Mathur, R.; Ho, J.E.; Pineda, P.; et al. Long-term evaluation of AAV-mediated sFlt-1 gene therapy for ocular neovascularization in mice and monkeys. Mol. Ther. 2005, 12, 659–668. [Google Scholar] [CrossRef]
  48. Li, Z.; He, T.; Du, K.; Xing, Y.Q.; Run, Y.M.; Yan, Y.; Shen, Y. Inhibition of oxygen-induced ischemic retinal neovascularization with adenoviral 15-lipoxygenase-1 gene transfer via up-regulation of PPAR-gamma and down-regulation of VEGFR-2 expression. PLoS ONE 2014, 9, e85824. [Google Scholar] [CrossRef]
  49. Qazi, Y.; Stagg, B.; Singh, N.; Singh, S.; Zhang, X.; Luo, L.; Simonis, J.; Kompella, U.B.; Ambati, B.K. Nanoparticle-mediated delivery of shRNA.VEGF-a plasmids regresses corneal neovascularization. Investig. Ophthalmol. Vis. Sci. 2012, 53, 2837–2844. [Google Scholar] [CrossRef]
  50. Torrecilla, J.; Gomez-Aguado, I.; Vicente-Pascual, M.; Del Pozo-Rodriguez, A.; Solinis, M.A.; Rodriguez-Gascon, A. MMP-9 Downregulation with Lipid Nanoparticles for Inhibiting Corneal Neovascularization by Gene Silencing. Nanomaterials 2019, 9, 631. [Google Scholar] [CrossRef]
  51. Yu, H.; Chen, L.; Jiang, J. Administration of pigment epithelium-derived factor delivered by adeno-associated virus inhibits blood-retinal barrier breakdown in diabetic rats. Mol. Vis. 2010, 16, 2384–2394. [Google Scholar] [PubMed]
  52. Coster, D.J.; Williams, K.A. The impact of corneal allograft rejection on the long-term outcome of corneal transplantation. Am. J. Ophthalmol. 2005, 140, 1112–1122. [Google Scholar] [CrossRef] [PubMed]
  53. Parker, D.G.; Brereton, H.M.; Coster, D.J.; Williams, K.A. The potential of viral vector-mediated gene transfer to prolong corneal allograft survival. Curr. Gene Ther. 2009, 9, 33–44. [Google Scholar] [CrossRef] [PubMed]
  54. Beutelspacher, S.C.; Pillai, R.; Watson, M.P.; Tan, P.H.; Tsang, J.; McClure, M.O.; George, A.J.; Larkin, D.F. Function of indoleamine 2,3-dioxygenase in corneal allograft rejection and prolongation of allograft survival by over-expression. Eur. J. Immunol. 2006, 36, 690–700. [Google Scholar] [CrossRef] [PubMed]
  55. Gong, N.; Pleyer, U.; Volk, H.D.; Ritter, T. Effects of local and systemic viral interleukin-10 gene transfer on corneal allograft survival. Gene Ther. 2007, 14, 484–490. [Google Scholar] [CrossRef] [PubMed]
  56. Gong, N.; Pleyer, U.; Yang, J.; Vogt, K.; Hill, M.; Anegon, I.; Volk, H.D.; Ritter, T. Influence of local and systemic CTLA4Ig gene transfer on corneal allograft survival. J. Gene Med. 2006, 8, 459–467. [Google Scholar] [CrossRef]
  57. Pillai, R.G.; Beutelspacher, S.C.; Larkin, D.F.; George, A.J. Expression of the chemokine antagonist vMIP II using a non-viral vector can prolong corneal allograft survival. Transplantation 2008, 85, 1640–1647. [Google Scholar] [CrossRef]
  58. Williams, K.A.; Coster, D.J. Gene therapy for diseases of the cornea—A review. Clin. Exp. Ophthalmol. 2010, 38, 93–103. [Google Scholar] [CrossRef]
  59. Murthy, R.C.; McFarland, T.J.; Yoken, J.; Chen, S.; Barone, C.; Burke, D.; Zhang, Y.; Appukuttan, B.; Stout, J.T. Corneal transduction to inhibit angiogenesis and graft failure. Investig. Ophthalmol. Vis. Sci. 2003, 44, 1837–1842. [Google Scholar] [CrossRef]
  60. Comer, R.M.; King, W.J.; Ardjomand, N.; Theoharis, S.; George, A.J.; Larkin, D.F. Effect of administration of CTLA4-Ig as protein or cDNA on corneal allograft survival. Investig. Ophthalmol. Vis. Sci. 2002, 43, 1095–1103. [Google Scholar]
  61. Tang, X.L.; Sun, J.F.; Wang, X.Y.; Du, L.L.; Liu, P. Blocking neuropilin-2 enhances corneal allograft survival by selectively inhibiting lymphangiogenesis on vascularized beds. Mol. Vis. 2010, 16, 2354–2361. [Google Scholar] [PubMed]
  62. Cho, Y.K.; Zhang, X.; Uehara, H.; Young, J.R.; Archer, B.; Ambati, B. Vascular Endothelial Growth Factor Receptor 1 morpholino increases graft survival in a murine penetrating keratoplasty model. Investig. Ophthalmol. Vis. Sci. 2012, 53, 8458–8471. [Google Scholar] [CrossRef] [PubMed]
  63. Swierkowska, J.; Gajecka, M. Genetic factors influencing the reduction of central corneal thickness in disorders affecting the eye. Ophthalmic Genet. 2017, 38, 501–510. [Google Scholar] [CrossRef]
  64. Dudakova, L.; Sasaki, T.; Liskova, P.; Palos, M.; Jirsova, K. The presence of lysyl oxidase-like enzymes in human control and keratoconic corneas. Histol. Histopathol. 2016, 31, 63–71. [Google Scholar] [CrossRef]
  65. Shetty, R.; Sathyanarayanamoorthy, A.; Ramachandra, R.A.; Arora, V.; Ghosh, A.; Srivatsa, P.R.; Pahuja, N.; Nuijts, R.M.; Sinha-Roy, A.; Mohan, R.R.; et al. Attenuation of lysyl oxidase and collagen gene expression in keratoconus patient corneal epithelium corresponds to disease severity. Mol. Vis. 2015, 21, 12–25. [Google Scholar]
  66. Smith, V.A.; Matthews, F.J.; Majid, M.A.; Cook, S.D. Keratoconus: Matrix metalloproteinase-2 activation and TIMP modulation. Biochim. Biophys. Acta 2006, 1762, 431–439. [Google Scholar] [CrossRef] [PubMed]
  67. Shetty, R.; Ghosh, A.; Lim, R.R.; Subramani, M.; Mihir, K.; Reshma, A.R.; Ranganath, A.; Nagaraj, S.; Nuijts, R.M.; Beuerman, R.; et al. Elevated expression of matrix metalloproteinase-9 and inflammatory cytokines in keratoconus patients is inhibited by cyclosporine A. Investig. Ophthalmol. Vis. Sci. 2015, 56, 738–750. [Google Scholar] [CrossRef] [PubMed]
  68. Shetty, R.; D’Souza, S.; Khamar, P.; Ghosh, A.; Nuijts, R.; Sethu, S. Biochemical Markers and Alterations in Keratoconus. Asia Pac. J. Ophthalmol. 2020, 9, 533–540. [Google Scholar] [CrossRef]
  69. Brancati, F.; Valente, E.M.; Sarkozy, A.; Feher, J.; Castori, M.; Del Duca, P.; Mingarelli, R.; Pizzuti, A.; Dallapiccola, B. A locus for autosomal dominant keratoconus maps to human chromosome 3p14-q13. J. Med. Genet. 2004, 41, 188–192. [Google Scholar] [CrossRef]
  70. Burdon, K.P.; Coster, D.J.; Charlesworth, J.C.; Mills, R.A.; Laurie, K.J.; Giunta, C.; Hewitt, A.W.; Latimer, P.; Craig, J.E. Apparent autosomal dominant keratoconus in a large Australian pedigree accounted for by digenic inheritance of two novel loci. Hum. Genet. 2008, 124, 379–386. [Google Scholar] [CrossRef]
  71. Hameed, A.; Khaliq, S.; Ismail, M.; Anwar, K.; Ebenezer, N.D.; Jordan, T.; Mehdi, S.Q.; Payne, A.M.; Bhattacharya, S.S. A novel locus for Leber congenital amaurosis (LCA4) with anterior keratoconus mapping to chromosome 17p13. Investig. Ophthalmol. Vis. Sci. 2000, 41, 629–633. [Google Scholar]
  72. Hutchings, H.; Ginisty, H.; Le Gallo, M.; Levy, D.; Stoesser, F.; Rouland, J.F.; Arne, J.L.; Lalaux, M.H.; Calvas, P.; Roth, M.P.; et al. Identification of a new locus for isolated familial keratoconus at 2p24. J. Med. Genet. 2005, 42, 88–94. [Google Scholar] [CrossRef] [PubMed]
  73. Wheeler, J.; Hauser, M.A.; Afshari, N.A.; Allingham, R.R.; Liu, Y. The Genetics of Keratoconus: A Review. Reprod. Syst. Sex Disord. 2012. [Google Scholar] [CrossRef]
  74. Czugala, M.; Karolak, J.A.; Nowak, D.M.; Polakowski, P.; Pitarque, J.; Molinari, A.; Rydzanicz, M.; Bejjani, B.A.; Yue, B.Y.; Szaflik, J.P.; et al. Novel mutation and three other sequence variants segregating with phenotype at keratoconus 13q32 susceptibility locus. Eur. J. Hum. Genet. 2012, 20, 389–397. [Google Scholar] [CrossRef] [PubMed]
  75. Droitcourt, C.; Touboul, D.; Ged, C.; Ezzedine, K.; Cario-Andre, M.; de Verneuil, H.; Colin, J.; Taieb, A. A prospective study of filaggrin null mutations in keratoconus patients with or without atopic disorders. Dermatology 2011, 222, 336–341. [Google Scholar] [CrossRef] [PubMed]
  76. Guan, T.; Liu, C.; Ma, Z.; Ding, S. The point mutation and polymorphism in keratoconus candidate gene TGFBI in Chinese population. Gene 2012, 503, 137–139. [Google Scholar] [CrossRef]
  77. Heon, E.; Greenberg, A.; Kopp, K.K.; Rootman, D.; Vincent, A.L.; Billingsley, G.; Priston, M.; Dorval, K.M.; Chow, R.L.; McInnes, R.R.; et al. VSX1: A gene for posterior polymorphous dystrophy and keratoconus. Hum. Mol. Genet. 2002, 11, 1029–1036. [Google Scholar] [CrossRef]
  78. Lechner, J.; Dash, D.P.; Muszynska, D.; Hosseini, M.; Segev, F.; George, S.; Frazer, D.G.; Moore, J.E.; Kaye, S.B.; Young, T.; et al. Mutational spectrum of the ZEB1 gene in corneal dystrophies supports a genotype-phenotype correlation. Investig. Ophthalmol. Vis. Sci. 2013, 54, 3215–3223. [Google Scholar] [CrossRef]
  79. Udar, N.; Atilano, S.R.; Brown, D.J.; Holguin, B.; Small, K.; Nesburn, A.B.; Kenney, M.C. SOD1: A candidate gene for keratoconus. Investig. Ophthalmol. Vis. Sci. 2006, 47, 3345–3351. [Google Scholar] [CrossRef]
  80. Gupta, N.; Vashist, P.; Ganger, A.; Tandon, R.; Gupta, S.K. Eye donation and eye banking in India. Natl. Med. J. India 2018, 31, 283–286. [Google Scholar] [CrossRef]
  81. Weiss, J.S.; Moller, H.U.; Aldave, A.J.; Seitz, B.; Bredrup, C.; Kivela, T.; Munier, F.L.; Rapuano, C.J.; Nischal, K.K.; Kim, E.K.; et al. IC3D classification of corneal dystrophies--edition 2. Cornea 2015, 34, 117–159. [Google Scholar] [CrossRef] [PubMed]
  82. Bourges, J.L. Corneal dystrophies. J. Fr. Ophtalmol. 2017, 40, e177–e192. [Google Scholar] [CrossRef]
  83. Soh, Y.Q.; Kocaba, V.; Weiss, J.S.; Jurkunas, U.V.; Kinoshita, S.; Aldave, A.J.; Mehta, J.S. Corneal dystrophies. Nat. Rev. Dis. Primers 2020, 6, 46. [Google Scholar] [CrossRef] [PubMed]
  84. Hamill, C.E.; Schmedt, T.; Jurkunas, U. Fuchs endothelial cornea dystrophy: A review of the genetics behind disease development. Semin. Ophthalmol. 2013, 28, 281–286. [Google Scholar] [CrossRef] [PubMed]
  85. Adamis, A.P.; Filatov, V.; Tripathi, B.J.; Tripathi, R.C. Fuchs’ endothelial dystrophy of the cornea. Surv. Ophthalmol. 1993, 38, 149–168. [Google Scholar] [CrossRef]
  86. Nishtala, K.; Pahuja, N.; Shetty, R.; Nuijts, R.M.; Ghosh, A. Tear biomarkers for keratoconus. Eye Vis. 2016, 3, 19. [Google Scholar] [CrossRef] [PubMed]
  87. Fan Gaskin, J.C.; Patel, D.V.; McGhee, C.N. Acute corneal hydrops in keratoconus—New perspectives. Am. J. Ophthalmol. 2014, 157, 921–928. [Google Scholar] [CrossRef]
  88. Hassan, O.M.; Farooq, A.V.; Soin, K.; Djalilian, A.R.; Hou, J.H. Management of Corneal Scarring Secondary to Herpes Zoster Keratitis. Cornea 2017, 36, 1018–1023. [Google Scholar] [CrossRef] [PubMed]
  89. Menda, S.A.; Das, M.; Panigrahi, A.; Prajna, N.V.; Acharya, N.R.; Lietman, T.M.; McLeod, S.D.; Keenan, J.D. Association of Postfungal Keratitis Corneal Scar Features with Visual Acuity. JAMA Ophthalmol. 2020, 138, 113–118. [Google Scholar] [CrossRef]
  90. Maharana, P.K.; Sharma, N.; Vajpayee, R.B. Acute corneal hydrops in keratoconus. Indian J. Ophthalmol. 2013, 61, 461–464. [Google Scholar] [CrossRef]
  91. Hashemi, H.; Heydarian, S.; Hooshmand, E.; Saatchi, M.; Yekta, A.; Aghamirsalim, M.; Valadkhan, M.; Mortazavi, M.; Hashemi, A.; Khabazkhoob, M. The Prevalence and Risk Factors for Keratoconus: A Systematic Review and Meta-Analysis. Cornea 2020, 39, 263–270. [Google Scholar] [CrossRef] [PubMed]
  92. Rebenitsch, R.L.; Kymes, S.M.; Walline, J.J.; Gordon, M.O. The lifetime economic burden of keratoconus: A decision analysis using a markov model. Am. J. Ophthalmol. 2011, 151, 768–773. [Google Scholar] [CrossRef] [PubMed]
  93. Tan, D.T.; Ang, L.P. Automated lamellar therapeutic keratoplasty for post-PRK corneal scarring and thinning. Am. J. Ophthalmol. 2004, 138, 1067–1069. [Google Scholar] [CrossRef] [PubMed]
  94. Deshmukh, R.; Reddy, J.C.; Rapuano, C.J.; Vaddavalli, P.K. Phototherapeutic keratectomy: Indications, methods and decision making. Indian J. Ophthalmol. 2020, 68, 2856–2866. [Google Scholar] [CrossRef] [PubMed]
  95. Tan, D.T.; Mehta, J.S. Future directions in lamellar corneal transplantation. Cornea 2007, 26, S21–S28. [Google Scholar] [CrossRef]
  96. Saini, J.S. Realistic targets and strategies in eye banking. Indian J. Ophthalmol. 1997, 45, 141–142. [Google Scholar]
  97. Dandona, L.; Naduvilath, T.J.; Janarthanan, M.; Ragu, K.; Rao, G.N. Survival analysis and visual outcome in a large series of corneal transplants in India. Br. J. Ophthalmol. 1997, 81, 726–731. [Google Scholar] [CrossRef]
  98. Sinha, R.; Vanathi, M.; Sharma, N.; Titiyal, J.S.; Vajpayee, R.B.; Tandon, R. Outcome of penetrating keratoplasty in patients with bilateral corneal blindness. Eye 2005, 19, 451–454. [Google Scholar] [CrossRef]
  99. Hos, D.; Matthaei, M.; Bock, F.; Maruyama, K.; Notara, M.; Clahsen, T.; Hou, Y.; Le, V.N.H.; Salabarria, A.C.; Horstmann, J.; et al. Immune reactions after modern lamellar (DALK, DSAEK, DMEK) versus conventional penetrating corneal transplantation. Prog. Retin. Eye Res. 2019, 73, 100768. [Google Scholar] [CrossRef]
  100. Alio Del Barrio, J.L.; Bhogal, M.; Ang, M.; Ziaei, M.; Robbie, S.; Montesel, A.; Gore, D.M.; Mehta, J.S.; Alio, J.L. Corneal transplantation after failed grafts: Options and outcomes. Surv. Ophthalmol. 2021, 66, 20–40. [Google Scholar] [CrossRef]
  101. Di Zazzo, A.; Kheirkhah, A.; Abud, T.B.; Goyal, S.; Dana, R. Management of high-risk corneal transplantation. Surv. Ophthalmol. 2017, 62, 816–827. [Google Scholar] [CrossRef] [PubMed]
  102. Tan, D.T.; Dart, J.K.; Holland, E.J.; Kinoshita, S. Corneal transplantation. Lancet 2012, 379, 1749–1761. [Google Scholar] [CrossRef] [PubMed]
  103. de By, T.M. Shortage in the face of plenty: Improving the allocation of corneas for transplantation. Dev. Ophthalmol. 2003, 36, 56–61. [Google Scholar] [CrossRef] [PubMed]
  104. Atchison, R.W.; Casto, B.C.; Hammon, W.M. Adenovirus-Associated Defective Virus Particles. Science 1965, 149, 754–756. [Google Scholar] [CrossRef] [PubMed]
  105. O’Donnell, J.; Taylor, K.A.; Chapman, M.S. Adeno-associated virus-2 and its primary cellular receptor--Cryo-EM structure of a heparin complex. Virology 2009, 385, 434–443. [Google Scholar] [CrossRef]
  106. Qing, K.; Mah, C.; Hansen, J.; Zhou, S.; Dwarki, V.; Srivastava, A. Human fibroblast growth factor receptor 1 is a co-receptor for infection by adeno-associated virus 2. Nat. Med. 1999, 5, 71–77. [Google Scholar] [CrossRef]
  107. Summerford, C.; Bartlett, J.S.; Samulski, R.J. AlphaVbeta5 integrin: A co-receptor for adeno-associated virus type 2 infection. Nat. Med. 1999, 5, 78–82. [Google Scholar] [CrossRef]
  108. Bartlett, J.S.; Wilcher, R.; Samulski, R.J. Infectious entry pathway of adeno-associated virus and adeno-associated virus vectors. J. Virol. 2000, 74, 2777–2785. [Google Scholar] [CrossRef]
  109. Douar, A.M.; Poulard, K.; Stockholm, D.; Danos, O. Intracellular trafficking of adeno-associated virus vectors: Routing to the late endosomal compartment and proteasome degradation. J. Virol. 2001, 75, 1824–1833. [Google Scholar] [CrossRef]
  110. Mohan, R.R.; Schultz, G.S.; Hong, J.W.; Mohan, R.R.; Wilson, S.E. Gene transfer into rabbit keratocytes using AAV and lipid-mediated plasmid DNA vectors with a lamellar flap for stromal access. Exp. Eye Res. 2003, 76, 373–383. [Google Scholar] [CrossRef]
  111. Chiorini, J.A.; Kim, F.; Yang, L.; Kotin, R.M. Cloning and characterization of adeno-associated virus type 5. J. Virol. 1999, 73, 1309–1319. [Google Scholar] [CrossRef] [PubMed]
  112. Zabner, J.; Seiler, M.; Walters, R.; Kotin, R.M.; Fulgeras, W.; Davidson, B.L.; Chiorini, J.A. Adeno-associated virus type 5 (AAV5) but not AAV2 binds to the apical surfaces of airway epithelia and facilitates gene transfer. J. Virol. 2000, 74, 3852–3858. [Google Scholar] [CrossRef]
  113. Gupta, S.; Rodier, J.T.; Sharma, A.; Giuliano, E.A.; Sinha, P.R.; Hesemann, N.P.; Ghosh, A.; Mohan, R.R. Targeted AAV5-Smad7 gene therapy inhibits corneal scarring in vivo. PLoS ONE 2017, 12, e0172928. [Google Scholar] [CrossRef] [PubMed]
  114. Mohan, R.R.; Tandon, A.; Sharma, A.; Cowden, J.W.; Tovey, J.C. Significant inhibition of corneal scarring in vivo with tissue-selective, targeted AAV5 decorin gene therapy. Investig. Ophthalmol. Vis. Sci. 2011, 52, 4833–4841. [Google Scholar] [CrossRef] [PubMed]
  115. Hippert, C.; Ibanes, S.; Serratrice, N.; Court, F.; Malecaze, F.; Kremer, E.J.; Kalatzis, V. Corneal transduction by intra-stromal injection of AAV vectors in vivo in the mouse and ex vivo in human explants. PLoS ONE 2012, 7, e35318. [Google Scholar] [CrossRef]
  116. Sharma, A.; Tovey, J.C.; Ghosh, A.; Mohan, R.R. AAV serotype influences gene transfer in corneal stroma in vivo. Exp. Eye Res. 2010, 91, 440–448. [Google Scholar] [CrossRef] [PubMed]
  117. Vance, M.; Llanga, T.; Bennett, W.; Woodard, K.; Murlidharan, G.; Chungfat, N.; Asokan, A.; Gilger, B.; Kurtzberg, J.; Samulski, R.J.; et al. AAV Gene Therapy for MPS1-associated Corneal Blindness. Sci. Rep. 2016, 6, 22131. [Google Scholar] [CrossRef] [PubMed]
  118. O’Callaghan, J.; Crosbie, D.E.; Cassidy, P.S.; Sherwood, J.M.; Flugel-Koch, C.; Lutjen-Drecoll, E.; Humphries, M.M.; Reina-Torres, E.; Wallace, D.; Kiang, A.S.; et al. Therapeutic potential of AAV-mediated MMP-3 secretion from corneal endothelium in treating glaucoma. Hum. Mol. Genet. 2017, 26, 1230–1246. [Google Scholar] [CrossRef]
  119. Bastola, P.; Song, L.; Gilger, B.C.; Hirsch, M.L. Adeno-Associated Virus Mediated Gene Therapy for Corneal Diseases. Pharmaceutics 2020, 12, 767. [Google Scholar] [CrossRef]
  120. Miyadera, K.; Conatser, L.; Llanga, T.A.; Carlin, K.; O’Donnell, P.; Bagel, J.; Song, L.; Kurtzberg, J.; Samulski, R.J.; Gilger, B.; et al. Intrastromal Gene Therapy Prevents and Reverses Advanced Corneal Clouding in a Canine Model of Mucopolysaccharidosis I. Mol. Ther. 2020, 28, 1455–1463. [Google Scholar] [CrossRef]
  121. Gilger, B.C.; Hirsch, M.L. Therapeutic Applications of Adeno-Associated Virus (AAV) Gene Transfer of HLA-G in the Eye. Int. J. Mol. Sci. 2022, 23, 3465. [Google Scholar] [CrossRef] [PubMed]
  122. Cheng, H.C.; Yeh, S.I.; Tsao, Y.P.; Kuo, P.C. Subconjunctival injection of recombinant AAV-angiostatin ameliorates alkali burn induced corneal angiogenesis. Mol. Vis. 2007, 13, 2344–2352. [Google Scholar] [PubMed]
  123. Lai, Y.K.; Shen, W.Y.; Brankov, M.; Lai, C.M.; Constable, I.J.; Rakoczy, P.E. Potential long-term inhibition of ocular neovascularisation by recombinant adeno-associated virus-mediated secretion gene therapy. Gene Ther. 2002, 9, 804–813. [Google Scholar] [CrossRef] [PubMed]
  124. Lu, Y.; Tai, P.W.L.; Ai, J.; Gessler, D.J.; Su, Q.; Yao, X.; Zheng, Q.; Zamore, P.D.; Xu, X.; Gao, G. Transcriptome Profiling of Neovascularized Corneas Reveals miR-204 as a Multi-target Biotherapy Deliverable by rAAVs. Mol. Ther. Nucleic Acids 2018, 10, 349–360. [Google Scholar] [CrossRef]
  125. Gupta, S.; Fink, M.K.; Kempuraj, D.; Sinha, N.R.; Martin, L.M.; Keele, L.M.; Sinha, P.R.; Giuliano, E.A.; Hesemann, N.P.; Raikwar, S.P.; et al. Corneal fibrosis abrogation by a localized AAV-mediated inhibitor of differentiation 3 (Id3) gene therapy in rabbit eyes in vivo. Mol. Ther. 2022, 30, 3257–3269. [Google Scholar] [CrossRef]
  126. Maddon, P.J.; McDougal, J.S.; Clapham, P.R.; Dalgleish, A.G.; Jamal, S.; Weiss, R.A.; Axel, R. HIV infection does not require endocytosis of its receptor, CD4. Cell 1988, 54, 865–874. [Google Scholar] [CrossRef]
  127. Gonda, M.A.; Wong-Staal, F.; Gallo, R.C.; Clements, J.E.; Narayan, O.; Gilden, R.V. Sequence homology and morphologic similarity of HTLV-III and visna virus, a pathogenic lentivirus. Science 1985, 227, 173–177. [Google Scholar] [CrossRef]
  128. Mohan, R.R.; Martin, L.M.; Sinha, N.R. Novel insights into gene therapy in the cornea. Exp. Eye Res. 2021, 202, 108361. [Google Scholar] [CrossRef]
  129. Mohan, R.R.; Tovey, J.C.; Sharma, A.; Schultz, G.S.; Cowden, J.W.; Tandon, A. Targeted decorin gene therapy delivered with adeno-associated virus effectively retards corneal neovascularization in vivo. PLoS ONE 2011, 6, e26432. [Google Scholar] [CrossRef]
  130. Barcia, R.N.; Dana, M.R.; Kazlauskas, A. Corneal graft rejection is accompanied by apoptosis of the endothelium and is prevented by gene therapy with bcl-xL. Am. J. Transplant. 2007, 7, 2082–2089. [Google Scholar] [CrossRef]
  131. Nosov, M.; Wilk, M.; Morcos, M.; Cregg, M.; O’Flynn, L.; Treacy, O.; Ritter, T. Role of lentivirus-mediated overexpression of programmed death-ligand 1 on corneal allograft survival. Am. J. Transplant. 2012, 12, 1313–1322. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, T.; Zhou, X.T.; Yu, Y.; Zhu, J.Y.; Dai, J.H.; Qu, X.M.; Le, Q.H.; Chu, R.Y. Inhibition of corneal fibrosis by Smad7 in rats after photorefractive keratectomy. Chin. Med. J. (Engl.) 2013, 126, 1445–1450. [Google Scholar] [PubMed]
  133. Parker, D.G.; Kaufmann, C.; Brereton, H.M.; Anson, D.S.; Francis-Staite, L.; Jessup, C.F.; Marshall, K.; Tan, C.; Koldej, R.; Coster, D.J.; et al. Lentivirus-mediated gene transfer to the rat, ovine and human cornea. Gene Ther. 2007, 14, 760–767. [Google Scholar] [CrossRef] [PubMed]
  134. Pastak, M.; Kleff, V.; Saban, D.R.; Czugala, M.; Steuhl, K.P.; Ergun, S.; Singer, B.B.; Fuchsluger, T.A. Gene Therapy for Modulation of T-Cell-Mediated Immune Response Provoked by Corneal Transplantation. Hum. Gene Ther. 2018, 29, 467–479. [Google Scholar] [CrossRef] [PubMed]
  135. Petrie, N.C.; Yao, F.; Eriksson, E. Gene therapy in wound healing. Surg. Clin. N. Am. 2003, 83, 597–616. [Google Scholar] [CrossRef]
  136. Borras, T.; Gabelt, B.T.; Klintworth, G.K.; Peterson, J.C.; Kaufman, P.L. Non-invasive observation of repeated adenoviral GFP gene delivery to the anterior segment of the monkey eye in vivo. J. Gene Med. 2001, 3, 437–449. [Google Scholar] [CrossRef]
  137. Fisher, K.D.; Stallwood, Y.; Green, N.K.; Ulbrich, K.; Mautner, V.; Seymour, L.W. Polymer-coated adenovirus permits efficient retargeting and evades neutralising antibodies. Gene Ther. 2001, 8, 341–348. [Google Scholar] [CrossRef]
  138. Saika, S.; Ikeda, K.; Yamanaka, O.; Flanders, K.C.; Nakajima, Y.; Miyamoto, T.; Ohnishi, Y.; Kao, W.W.; Muragaki, Y.; Ooshima, A. Therapeutic effects of adenoviral gene transfer of bone morphogenic protein-7 on a corneal alkali injury model in mice. Lab. Investig. 2005, 85, 474–486. [Google Scholar] [CrossRef]
  139. Nguyen, P.; Yiu, S.C. Strategies for local gene therapy of corneal allograft rejection. Middle East Afr. J. Ophthalmol. 2013, 20, 11–25. [Google Scholar] [CrossRef]
  140. Zhou, S.Y.; Xie, Z.L.; Xiao, O.; Yang, X.R.; Heng, B.C.; Sato, Y. Inhibition of mouse alkali burn induced-corneal neovascularization by recombinant adenovirus encoding human vasohibin-1. Mol. Vis. 2010, 16, 1389–1398. [Google Scholar]
  141. Yu, H.; Wu, J.; Li, H.; Wang, Z.; Chen, X.; Tian, Y.; Yi, M.; Ji, X.; Ma, J.; Huang, Q. Inhibition of corneal neovascularization by recombinant adenovirus-mediated sFlk-1 expression. Biochem. Biophys. Res. Commun. 2007, 361, 946–952. [Google Scholar] [CrossRef] [PubMed]
  142. Saika, S.; Yamanaka, O.; Okada, Y.; Miyamoto, T.; Kitano, A.; Flanders, K.C.; Ohnishi, Y.; Nakajima, Y.; Kao, W.W.; Ikeda, K. Effect of overexpression of PPARgamma on the healing process of corneal alkali burn in mice. Am. J. Physiol. Cell Physiol. 2007, 293, C75–C86. [Google Scholar] [CrossRef] [PubMed]
  143. Saghizadeh, M.; Kramerov, A.A.; Yu, F.S.; Castro, M.G.; Ljubimov, A.V. Normalization of wound healing and diabetic markers in organ cultured human diabetic corneas by adenoviral delivery of c-Met gene. Investig. Ophthalmol. Vis. Sci. 2010, 51, 1970–1980. [Google Scholar] [CrossRef] [PubMed]
  144. Klebe, S.; Sykes, P.J.; Coster, D.J.; Krishnan, R.; Williams, K.A. Prolongation of sheep corneal allograft survival by ex vivo transfer of the gene encoding interleukin-10. Transplantation 2001, 71, 1214–1220. [Google Scholar] [CrossRef]
  145. Klebe, S.; Coster, D.J.; Sykes, P.J.; Swinburne, S.; Hallsworth, P.; Scheerlinck, J.P.; Krishnan, R.; Williams, K.A. Prolongation of sheep corneal allograft survival by transfer of the gene encoding ovine IL-12-p40 but not IL-4 to donor corneal endothelium. J. Immunol. 2005, 175, 2219–2226. [Google Scholar] [CrossRef]
  146. Trapani, I.; Puppo, A.; Auricchio, A. Vector platforms for gene therapy of inherited retinopathies. Prog. Retin. Eye Res. 2014, 43, 108–128. [Google Scholar] [CrossRef]
  147. Chang, Y.K.; Hwang, J.S.; Chung, T.Y.; Shin, Y.J. SOX2 Activation Using CRISPR/dCas9 Promotes Wound Healing in Corneal Endothelial Cells. Stem Cells 2018, 36, 1851–1862. [Google Scholar] [CrossRef]
  148. Zhou, R.; Dean, D.A. Gene transfer of interleukin 10 to the murine cornea using electroporation. Exp. Biol. Med. (Maywood) 2007, 232, 362–369. [Google Scholar]
  149. Oshima, Y.; Sakamoto, T.; Hisatomi, T.; Tsutsumi, C.; Sassa, Y.; Ishibashi, T.; Inomata, H. Targeted gene transfer to corneal stroma in vivo by electric pulses. Exp. Eye Res. 2002, 74, 191–198. [Google Scholar] [CrossRef]
  150. Rosazza, C.; Meglic, S.H.; Zumbusch, A.; Rols, M.P.; Miklavcic, D. Gene Electrotransfer: A Mechanistic Perspective. Curr. Gene Ther. 2016, 16, 98–129. [Google Scholar] [CrossRef]
  151. de la Fuente, M.; Seijo, B.; Alonso, M.J. Bioadhesive hyaluronan-chitosan nanoparticles can transport genes across the ocular mucosa and transfect ocular tissue. Gene Ther. 2008, 15, 668–676. [Google Scholar] [CrossRef] [PubMed]
  152. Jain, G.K.; Pathan, S.A.; Akhter, S.; Jayabalan, N.; Talegaonkar, S.; Khar, R.K.; Ahmad, F.J. Microscopic and spectroscopic evaluation of novel PLGA-chitosan Nanoplexes as an ocular delivery system. Colloids Surf. B Biointerfaces 2011, 82, 397–403. [Google Scholar] [CrossRef]
  153. Nagarwal, R.C.; Singh, P.N.; Kant, S.; Maiti, P.; Pandit, J.K. Chitosan nanoparticles of 5-fluorouracil for ophthalmic delivery: Characterization, in-vitro and in-vivo study. Chem. Pharm. Bull. 2011, 59, 272–278. [Google Scholar] [CrossRef] [PubMed]
  154. Felgner, P.L.; Gadek, T.R.; Holm, M.; Roman, R.; Chan, H.W.; Wenz, M.; Northrop, J.P.; Ringold, G.M.; Danielsen, M. Lipofection: A highly efficient, lipid-mediated DNA-transfection procedure. Proc. Natl. Acad. Sci. USA 1987, 84, 7413–7417. [Google Scholar] [CrossRef] [PubMed]
  155. Farhood, H.; Serbina, N.; Huang, L. The role of dioleoyl phosphatidylethanolamine in cationic liposome mediated gene transfer. Biochim. Biophys. Acta 1995, 1235, 289–295. [Google Scholar] [CrossRef]
  156. Taketani, Y.; Kitamoto, K.; Sakisaka, T.; Kimakura, M.; Toyono, T.; Yamagami, S.; Amano, S.; Kuroda, M.; Moore, T.; Usui, T.; et al. Repair of the TGFBI gene in human corneal keratocytes derived from a granular corneal dystrophy patient via CRISPR/Cas9-induced homology-directed repair. Sci. Rep. 2017, 7, 16713. [Google Scholar] [CrossRef]
  157. Courtney, D.G.; Atkinson, S.D.; Allen, E.H.; Moore, J.E.; Walsh, C.P.; Pedrioli, D.M.; MacEwen, C.J.; Pellegrini, G.; Maurizi, E.; Serafini, C.; et al. siRNA silencing of the mutant keratin 12 allele in corneal limbal epithelial cells grown from patients with Meesmann’s epithelial corneal dystrophy. Investig. Ophthalmol. Vis. Sci. 2014, 55, 3352–3360. [Google Scholar] [CrossRef]
  158. Courtney, D.G.; Atkinson, S.D.; Moore, J.E.; Maurizi, E.; Serafini, C.; Pellegrini, G.; Black, G.C.; Manson, F.D.; Yam, G.H.; Macewen, C.J.; et al. Development of allele-specific gene-silencing siRNAs for TGFBI Arg124Cys in lattice corneal dystrophy type I. Investig. Ophthalmol. Vis. Sci. 2014, 55, 977–985. [Google Scholar] [CrossRef] [PubMed]
  159. Donnelly, K.S.; Giuliano, E.A.; Sharma, A.; Tandon, A.; Rodier, J.T.; Mohan, R.R. Decorin-PEI nanoconstruct attenuates equine corneal fibroblast differentiation. Vet. Ophthalmol. 2014, 17, 162–169. [Google Scholar] [CrossRef]
  160. Hu, M.L.; Edwards, T.L.; O’Hare, F.; Hickey, D.G.; Wang, J.H.; Liu, Z.; Ayton, L.N. Gene therapy for inherited retinal diseases: Progress and possibilities. Clin. Exp. Optom. 2021, 104, 444–454. [Google Scholar] [CrossRef]
  161. Lee, J.H.; Wang, J.H.; Chen, J.; Li, F.; Edwards, T.L.; Hewitt, A.W.; Liu, G.S. Gene therapy for visual loss: Opportunities and concerns. Prog. Retin. Eye Res. 2019, 68, 31–53. [Google Scholar] [CrossRef] [PubMed]
  162. Chtarto, A.; Bockstael, O.; Gebara, E.; Vermoesen, K.; Melas, C.; Pythoud, C.; Levivier, M.; De Witte, O.; Luthi-Carter, R.; Clinkers, R.; et al. An adeno-associated virus-based intracellular sensor of pathological nuclear factor-kappaB activation for disease-inducible gene transfer. PLoS ONE 2013, 8, e53156. [Google Scholar] [CrossRef] [PubMed]
  163. Burnight, E.R.; Giacalone, J.C.; Cooke, J.A.; Thompson, J.R.; Bohrer, L.R.; Chirco, K.R.; Drack, A.V.; Fingert, J.H.; Worthington, K.S.; Wiley, L.A.; et al. CRISPR-Cas9 genome engineering: Treating inherited retinal degeneration. Prog. Retin. Eye Res. 2018, 65, 28–49. [Google Scholar] [CrossRef] [PubMed]
  164. Yin, H.; Kanasty, R.L.; Eltoukhy, A.A.; Vegas, A.J.; Dorkin, J.R.; Anderson, D.G. Non-viral vectors for gene-based therapy. Nat. Rev. Genet. 2014, 15, 541–555. [Google Scholar] [CrossRef] [PubMed]
  165. Raikwar, S.P.; Raikwar, A.S.; Chaurasia, S.S.; Mohan, R.R. Gene editing for corneal disease management. World J. Transl. Med. 2016, 5, 1–13. [Google Scholar] [CrossRef] [PubMed]
  166. Ma, E.; Harrington, L.B.; O’Connell, M.R.; Zhou, K.; Doudna, J.A. Single-Stranded DNA Cleavage by Divergent CRISPR-Cas9 Enzymes. Mol. Cell 2015, 60, 398–407. [Google Scholar] [CrossRef]
  167. Yu, W.; Wu, Z. Ocular delivery of CRISPR/Cas genome editing components for treatment of eye diseases. Adv. Drug Deliv. Rev. 2021, 168, 181–195. [Google Scholar] [CrossRef]
  168. Yang, Y.; Wang, L.; Bell, P.; McMenamin, D.; He, Z.; White, J.; Yu, H.; Xu, C.; Morizono, H.; Musunuru, K.; et al. A dual AAV system enables the Cas9-mediated correction of a metabolic liver disease in newborn mice. Nat. Biotechnol. 2016, 34, 334–338. [Google Scholar] [CrossRef]
  169. Courtney, D.G.; Moore, J.E.; Atkinson, S.D.; Maurizi, E.; Allen, E.H.; Pedrioli, D.M.; McLean, W.H.; Nesbit, M.A.; Moore, C.B. CRISPR/Cas9 DNA cleavage at SNP-derived PAM enables both in vitro and in vivo KRT12 mutation-specific targeting. Gene Ther. 2016, 23, 108–112. [Google Scholar] [CrossRef]
  170. Elbashir, S.M.; Harborth, J.; Lendeckel, W.; Yalcin, A.; Weber, K.; Tuschl, T. Duplexes of 21-nucleotide RNAs mediate RNA interference in cultured mammalian cells. Nature 2001, 411, 494–498. [Google Scholar] [CrossRef]
  171. Hammond, S.M.; Bernstein, E.; Beach, D.; Hannon, G.J. An RNA-directed nuclease mediates post-transcriptional gene silencing in Drosophila cells. Nature 2000, 404, 293–296. [Google Scholar] [CrossRef] [PubMed]
  172. Liu, Q.; Wu, K.; Qiu, X.; Yang, Y.; Lin, X.; Yu, M. siRNA silencing of gene expression in trabecular meshwork: RhoA siRNA reduces IOP in mice. Curr. Mol. Med. 2012, 12, 1015–1027. [Google Scholar] [CrossRef] [PubMed]
  173. Supe, S.; Upadhya, A.; Singh, K. Role of small interfering RNA (siRNA) in targeting ocular neovascularization: A review. Exp. Eye Res. 2021, 202, 108329. [Google Scholar] [CrossRef] [PubMed]
  174. Bennett, C.F.; Krainer, A.R.; Cleveland, D.W. Antisense Oligonucleotide Therapies for Neurodegenerative Diseases. Annu. Rev. Neurosci. 2019, 42, 385–406. [Google Scholar] [CrossRef]
  175. Cursiefen, C.; Viaud, E.; Bock, F.; Geudelin, B.; Ferry, A.; Kadlecova, P.; Levy, M.; Al Mahmood, S.; Colin, S.; Thorin, E.; et al. Aganirsen antisense oligonucleotide eye drops inhibit keratitis-induced corneal neovascularization and reduce need for transplantation: The I-CAN study. Ophthalmology 2014, 121, 1683–1692. [Google Scholar] [CrossRef]
  176. Cursiefen, C.; Bock, F.; Horn, F.K.; Kruse, F.E.; Seitz, B.; Borderie, V.; Fruh, B.; Thiel, M.A.; Wilhelm, F.; Geudelin, B.; et al. GS-101 antisense oligonucleotide eye drops inhibit corneal neovascularization: Interim results of a randomized phase II trial. Ophthalmology 2009, 116, 1630–1637. [Google Scholar] [CrossRef]
  177. Ghosh, A.; Duan, D. Expanding adeno-associated viral vector capacity: A tale of two vectors. Biotechnol. Genet. Eng. Rev. 2007, 24, 165–177. [Google Scholar] [CrossRef]
  178. Ghosh, A.; Yue, Y.; Duan, D. Efficient transgene reconstitution with hybrid dual AAV vectors carrying the minimized bridging sequences. Hum. Gene Ther. 2011, 22, 77–83. [Google Scholar] [CrossRef]
  179. Ginn, S.L.; Amaya, A.K.; Alexander, I.E.; Edelstein, M.; Abedi, M.R. Gene therapy clinical trials worldwide to 2017: An update. J. Gene Med. 2018, 20, e3015. [Google Scholar] [CrossRef]
  180. Di Iorio, E.; Barbaro, V.; Alvisi, G.; Trevisan, M.; Ferrari, S.; Masi, G.; Nespeca, P.; Ghassabian, H.; Ponzin, D.; Palu, G. New Frontiers of Corneal Gene Therapy. Hum. Gene Ther. 2019, 30, 923–945. [Google Scholar] [CrossRef]
  181. Mohan, R.R.; Tripathi, R.; Sharma, A.; Sinha, P.R.; Giuliano, E.A.; Hesemann, N.P.; Chaurasia, S.S. Decorin antagonizes corneal fibroblast migration via caveolae-mediated endocytosis of epidermal growth factor receptor. Exp. Eye Res. 2019, 180, 200–207. [Google Scholar] [CrossRef] [PubMed]
  182. Gupta, S.; Fink, M.K.; Ghosh, A.; Tripathi, R.; Sinha, P.R.; Sharma, A.; Hesemann, N.P.; Chaurasia, S.S.; Giuliano, E.A.; Mohan, R.R. Novel Combination BMP7 and HGF Gene Therapy Instigates Selective Myofibroblast Apoptosis and Reduces Corneal Haze In Vivo. Investig. Ophthalmol. Vis. Sci. 2018, 59, 1045–1057. [Google Scholar] [CrossRef] [PubMed]
  183. Chaudhary, K.; Moore, H.; Tandon, A.; Gupta, S.; Khanna, R.; Mohan, R.R. Nanotechnology and adeno-associated virus-based decorin gene therapy ameliorates peritoneal fibrosis. Am. J. Physiol. Renal. Physiol. 2014, 307, F777–F782. [Google Scholar] [CrossRef] [PubMed]
  184. Lai, L.J.; Xiao, X.; Wu, J.H. Inhibition of corneal neovascularization with endostatin delivered by adeno-associated viral (AAV) vector in a mouse corneal injury model. J. Biomed. Sci. 2007, 14, 313–322. [Google Scholar] [CrossRef] [PubMed]
  185. Parker, D.G.; Coster, D.J.; Brereton, H.M.; Hart, P.H.; Koldej, R.; Anson, D.S.; Williams, K.A. Lentivirus-mediated gene transfer of interleukin 10 to the ovine and human cornea. Clin. Exp. Ophthalmol. 2010, 38, 405–413. [Google Scholar] [CrossRef] [PubMed]
  186. Fuchsluger, T.A.; Jurkunas, U.; Kazlauskas, A.; Dana, R. Anti-apoptotic gene therapy prolongs survival of corneal endothelial cells during storage. Gene Ther. 2011, 18, 778–787. [Google Scholar] [CrossRef]
  187. Watson, Z.L.; Washington, S.D.; Phelan, D.M.; Lewin, A.S.; Tuli, S.S.; Schultz, G.S.; Neumann, D.M.; Bloom, D.C. In Vivo Knockdown of the Herpes Simplex Virus 1 Latency-Associated Transcript Reduces Reactivation from Latency. J. Virol. 2018, 92. [Google Scholar] [CrossRef]
  188. Elbadawy, H.M.; Gailledrat, M.; Desseaux, C.; Salvalaio, G.; Di Iorio, E.; Ferrari, B.; Bertolin, M.; Barbaro, V.; Parekh, M.; Gayon, R.; et al. Gene transfer of integration defective anti-HSV-1 meganuclease to human corneas ex vivo. Gene Ther. 2014, 21, 272–281. [Google Scholar] [CrossRef]
  189. Sawamoto, K.; Chen, H.H.; Almeciga-Diaz, C.J.; Mason, R.W.; Tomatsu, S. Gene therapy for Mucopolysaccharidoses. Mol. Genet. Metab. 2018, 123, 59–68. [Google Scholar] [CrossRef]
  190. Song, L.; Bower, J.J.; Llanga, T.; Salmon, J.H.; Hirsch, M.L.; Gilger, B.C. Ocular Tolerability and Immune Response to Corneal Intrastromal AAV-IDUA Gene Therapy in New Zealand White Rabbits. Mol. Ther. Methods Clin. Dev. 2020, 18, 24–32. [Google Scholar] [CrossRef]
  191. Kamata, Y.; Okuyama, T.; Kosuga, M.; O’Hira, A.; Kanaji, A.; Sasaki, K.; Yamada, M.; Azuma, N. Adenovirus-mediated gene therapy for corneal clouding in mice with mucopolysaccharidosis type VII. Mol. Ther. 2001, 4, 307–312. [Google Scholar] [CrossRef] [PubMed]
  192. Serratrice, N.; Cubizolle, A.; Ibanes, S.; Mestre-Frances, N.; Bayo-Puxan, N.; Creyssels, S.; Gennetier, A.; Bernex, F.; Verdier, J.M.; Haskins, M.E.; et al. Corrective GUSB transfer to the canine mucopolysaccharidosis VII cornea using a helper-dependent canine adenovirus vector. J. Control. Release 2014, 181, 22–31. [Google Scholar] [CrossRef] [PubMed]
  193. Cotugno, G.; Annunziata, P.; Tessitore, A.; O’Malley, T.; Capalbo, A.; Faella, A.; Bartolomeo, R.; O’Donnell, P.; Wang, P.; Russo, F.; et al. Long-term amelioration of feline Mucopolysaccharidosis VI after AAV-mediated liver gene transfer. Mol. Ther. 2011, 19, 461–469. [Google Scholar] [CrossRef] [PubMed]
  194. Carlson, E.C.; Liu, C.Y.; Yang, X.; Gregory, M.; Ksander, B.; Drazba, J.; Perez, V.L. In vivo gene delivery and visualization of corneal stromal cells using an adenoviral vector and keratocyte-specific promoter. Investig. Ophthalmol. Vis. Sci. 2004, 45, 2194–2200. [Google Scholar] [CrossRef]
  195. Liu, J.J.; Kao, W.W.; Wilson, S.E. Corneal epithelium-specific mouse keratin K12 promoter. Exp. Eye Res. 1999, 68, 295–301. [Google Scholar] [CrossRef]
  196. Hamilton, A.I. Keratoblast and keratocyte, not keratinocyte. Nature 1972, 238, 98. [Google Scholar] [CrossRef]
  197. Liu, C.; Arar, H.; Kao, C.; Kao, W.W. Identification of a 3.2 kb 5′-flanking region of the murine keratocan gene that directs beta-galactosidase expression in the adult corneal stroma of transgenic mice. Gene 2000, 250, 85–96. [Google Scholar] [CrossRef]
  198. Korecki, A.J.; Cueva-Vargas, J.L.; Fornes, O.; Agostinone, J.; Farkas, R.A.; Hickmott, J.W.; Lam, S.L.; Mathelier, A.; Zhou, M.; Wasserman, W.W.; et al. Human MiniPromoters for ocular-rAAV expression in ON bipolar, cone, corneal, endothelial, Muller glial, and PAX6 cells. Gene Ther. 2021, 28, 351–372. [Google Scholar] [CrossRef]
  199. Sakai, R.; Kinouchi, T.; Kawamoto, S.; Dana, M.R.; Hamamoto, T.; Tsuru, T.; Okubo, K.; Yamagami, S. Construction of human corneal endothelial cDNA library and identification of novel active genes. Investig. Ophthalmol. Vis. Sci. 2002, 43, 1749–1756. [Google Scholar]
  200. Verkman, A.S. Role of aquaporin water channels in eye function. Exp. Eye Res. 2003, 76, 137–143. [Google Scholar] [CrossRef]
  201. Johari, Y.B.; Mercer, A.C.; Liu, Y.; Brown, A.J.; James, D.C. Design of synthetic promoters for controlled expression of therapeutic genes in retinal pigment epithelial cells. Biotechnol. Bioeng. 2021, 118, 2001–2015. [Google Scholar] [CrossRef] [PubMed]
  202. Croze, R.H.; Kotterman, M.; Burns, C.H.; Schmitt, C.E.; Quezada, M.; Schaffer, D.; Kirn, D.; Francis, P. Viral Vector Technologies and Strategies: Improving on Nature. Int. Ophthalmol. Clin. 2021, 61, 59–89. [Google Scholar] [CrossRef] [PubMed]
  203. Lerch, T.F.; Xie, Q.; Chapman, M.S. The structure of adeno-associated virus serotype 3B (AAV-3B): Insights into receptor binding and immune evasion. Virology 2010, 403, 26–36. [Google Scholar] [CrossRef] [PubMed]
  204. Meyer, N.L.; Hu, G.; Davulcu, O.; Xie, Q.; Noble, A.J.; Yoshioka, C.; Gingerich, D.S.; Trzynka, A.; David, L.; Stagg, S.M.; et al. Structure of the gene therapy vector, adeno-associated virus with its cell receptor, AAVR. Elife 2019, 8, e44707. [Google Scholar] [CrossRef] [PubMed]
  205. Ng, R.; Govindasamy, L.; Gurda, B.L.; McKenna, R.; Kozyreva, O.G.; Samulski, R.J.; Parent, K.N.; Baker, T.S.; Agbandje-McKenna, M. Structural characterization of the dual glycan binding adeno-associated virus serotype 6. J. Virol. 2010, 84, 12945–12957. [Google Scholar] [CrossRef]
  206. Shen, S.; Troupes, A.N.; Pulicherla, N.; Asokan, A. Multiple roles for sialylated glycans in determining the cardiopulmonary tropism of adeno-associated virus 4. J. Virol. 2013, 87, 13206–13213. [Google Scholar] [CrossRef]
  207. Mays, L.E.; Wang, L.; Tenney, R.; Bell, P.; Nam, H.J.; Lin, J.; Gurda, B.; Van Vliet, K.; Mikals, K.; Agbandje-McKenna, M.; et al. Mapping the structural determinants responsible for enhanced T cell activation to the immunogenic adeno-associated virus capsid from isolate rhesus 32.33. J. Virol. 2013, 87, 9473–9485. [Google Scholar] [CrossRef]
  208. Gurda, B.L.; DiMattia, M.A.; Miller, E.B.; Bennett, A.; McKenna, R.; Weichert, W.S.; Nelson, C.D.; Chen, W.J.; Muzyczka, N.; Olson, N.H.; et al. Capsid antibodies to different adeno-associated virus serotypes bind common regions. J. Virol. 2013, 87, 9111–9124. [Google Scholar] [CrossRef]
  209. Tseng, Y.S.; Gurda, B.L.; Chipman, P.; McKenna, R.; Afione, S.; Chiorini, J.A.; Muzyczka, N.; Olson, N.H.; Baker, T.S.; Kleinschmidt, J.; et al. Adeno-associated virus serotype 1 (AAV1)- and AAV5-antibody complex structures reveal evolutionary commonalities in parvovirus antigenic reactivity. J. Virol. 2015, 89, 1794–1808. [Google Scholar] [CrossRef]
  210. Wobus, C.E.; Hugle-Dorr, B.; Girod, A.; Petersen, G.; Hallek, M.; Kleinschmidt, J.A. Monoclonal antibodies against the adeno-associated virus type 2 (AAV-2) capsid: Epitope mapping and identification of capsid domains involved in AAV-2-cell interaction and neutralization of AAV-2 infection. J. Virol. 2000, 74, 9281–9293. [Google Scholar] [CrossRef]
  211. Mah, C.; Qing, K.; Khuntirat, B.; Ponnazhagan, S.; Wang, X.S.; Kube, D.M.; Yoder, M.C.; Srivastava, A. Adeno-associated virus type 2-mediated gene transfer: Role of epidermal growth factor receptor protein tyrosine kinase in transgene expression. J. Virol. 1998, 72, 9835–9843. [Google Scholar] [CrossRef] [PubMed]
  212. Qing, K.; Wang, X.S.; Kube, D.M.; Ponnazhagan, S.; Bajpai, A.; Srivastava, A. Role of tyrosine phosphorylation of a cellular protein in adeno-associated virus 2-mediated transgene expression. Proc. Natl. Acad. Sci. USA 1997, 94, 10879–10884. [Google Scholar] [CrossRef] [PubMed]
  213. Zhong, L.; Zhao, W.; Wu, J.; Li, B.; Zolotukhin, S.; Govindasamy, L.; Agbandje-McKenna, M.; Srivastava, A. A dual role of EGFR protein tyrosine kinase signaling in ubiquitination of AAV2 capsids and viral second-strand DNA synthesis. Mol. Ther. 2007, 15, 1323–1330. [Google Scholar] [CrossRef] [PubMed]
  214. Markusic, D.M.; Herzog, R.W.; Aslanidi, G.V.; Hoffman, B.E.; Li, B.; Li, M.; Jayandharan, G.R.; Ling, C.; Zolotukhin, I.; Ma, W.; et al. High-efficiency transduction and correction of murine hemophilia B using AAV2 vectors devoid of multiple surface-exposed tyrosines. Mol. Ther. 2010, 18, 2048–2056. [Google Scholar] [CrossRef] [PubMed]
  215. Kay, C.N.; Ryals, R.C.; Aslanidi, G.V.; Min, S.H.; Ruan, Q.; Sun, J.; Dyka, F.M.; Kasuga, D.; Ayala, A.E.; Van Vliet, K.; et al. Targeting photoreceptors via intravitreal delivery using novel, capsid-mutated AAV vectors. PLoS ONE 2013, 8, e62097. [Google Scholar] [CrossRef]
Figure 1. Corneal dystrophies across various corneal layers. Corneal dystrophies affect different layers of the cornea including epithelium, Bowman’s layer, stroma, and Descemet’s membrane and endothelium. Created with Biorender.com (accessed on 14 March 2023).
Figure 1. Corneal dystrophies across various corneal layers. Corneal dystrophies affect different layers of the cornea including epithelium, Bowman’s layer, stroma, and Descemet’s membrane and endothelium. Created with Biorender.com (accessed on 14 March 2023).
Cells 12 01280 g001
Figure 2. Schematic illustration of (A) AAV genome, and (B) mechanism of entry, infection, and therapeutic gene production in host cell. Created with Biorender.com (accessed on 14 March 2023).
Figure 2. Schematic illustration of (A) AAV genome, and (B) mechanism of entry, infection, and therapeutic gene production in host cell. Created with Biorender.com (accessed on 14 March 2023).
Cells 12 01280 g002
Figure 3. Schematic illustration representing mechanism of action of (A) adenovirus and (B) lentivirus. Created with Biorender.com (accessed on 14 March 2023).
Figure 3. Schematic illustration representing mechanism of action of (A) adenovirus and (B) lentivirus. Created with Biorender.com (accessed on 14 March 2023).
Cells 12 01280 g003
Figure 4. Delivery mechanism for non-viral vectors. Created with Biorender.com (accessed on 14 March 2023).
Figure 4. Delivery mechanism for non-viral vectors. Created with Biorender.com (accessed on 14 March 2023).
Cells 12 01280 g004
Figure 5. Schematic illustration for the different strategies used in corneal gene therapy. (A) Gene augmentation. (B) Gene silencing using ASO, siRNA, and miRNA. (C) Gene editing using CRISPR Cas9 approach. Created with Biorender.com (accessed on 14 March 2023).
Figure 5. Schematic illustration for the different strategies used in corneal gene therapy. (A) Gene augmentation. (B) Gene silencing using ASO, siRNA, and miRNA. (C) Gene editing using CRISPR Cas9 approach. Created with Biorender.com (accessed on 14 March 2023).
Cells 12 01280 g005
Figure 6. Schematic illustration representing dual AAV approach for large genes. Created with Biorender.com (accessed on 14 March 2023).
Figure 6. Schematic illustration representing dual AAV approach for large genes. Created with Biorender.com (accessed on 14 March 2023).
Cells 12 01280 g006
Figure 7. Routes of administration for corneal gene therapy. Created with Biorender.com (accessed on 14 March 2023).
Figure 7. Routes of administration for corneal gene therapy. Created with Biorender.com (accessed on 14 March 2023).
Cells 12 01280 g007
Table 1. Classification of Corneal dystrophies.
Table 1. Classification of Corneal dystrophies.
RegionName of DystrophyGeneGene LocusMode of InheritanceIC3D CategoryAge of OnsetSymptomsVisual AcuityClinical Appearance of the Cornea
Epithelial and Subepithelial DystrophiesEpithelial basement membrane dystrophy (EBMD)TGFBI5q31Sporadic1AdultCorneal erosion, slightly distorted visionMild visual reductionThickening of the epithelium, round or oval opacities, and lines.
Epithelial recurrent erosion dystrophy (ERED)COL17E110q23Autosomal Dominant31st DecadeStinging, burning, painful corneal erosion, photophobiaSometimes impairedEpithelial erosion
Subepithelial mucinous corneal dystrophy (SMCD)UnknownunknownAutosomal Dominant41st DecadePainful incidence of recurrent corneal erosionProgressive loss of visionBilateral subepithelial opacities and haze, mostly denser centrally, involving the entire cornea
Meesmann corneal dystrophy (MECD)KRT3, KRT12,12q13, 17q12Autosomal Dominant1Early childhoodMild erosion and reduced sensation of the corneaRarely blurred visionMultiple, tiny, distinct epithelial vesicles extend to the limbus and are mostly cumulated in the interpalpebral area.
Lisch epithelial corneal dystrophy (LECD)UnknownXp22.3X-linked Dominant2Childhoodsymptomatic or blurred vision if the pupillary zone is involvedSometimes impairedLocalized epithelial opacities of various patterns: whorls, bands, flame, feather shaped.
Gelatinous drop-like corneal dystrophy (GDLD)TACSTD2 (M1S1)1p32Autosomal Recessive11st to 2nd DecadeDistorted vision, photophobia, scratchy sensation, redness, tearingMarked visual impairmentAppearance of subepithelial lesions, indicating extremely hyperpermeable corneal epithelium, Superficial vascularization, stromal opacification
Bowman Layer DystrophiesReis–Buckler’s corneal dystrophy (RBCD)TGFBI5q31Autosomal Dominant1ChildhoodPainful incidence of recurrent corneal erosionProgressive deuteriation of visionReplacement of bowman layer by sheet like connective tissue with granular deposits, which extends to subepithelial stroma.
Thiel–Behnke corneal dystrophy (TBCD)TGFBI5q31Autosomal Dominant2ChildhoodPainful incidence of recurrent corneal erosionGradual visual impairmentSymmetrical subepithelial reticular (honeycomb) opacities in the central cornea; which can progress to deep stromal layers and corneal periphery.
Stromal DystrophiesLattice corneal dystrophy (LCD)TGFBI5q31Autosomal Dominant11st DecadeStinging, burning, Painful incidence of recurrent corneal erosionProgressive visual impairmentThin branching refractile lines or subepithelial ovoid dots in the central cornea, diffuse stromal, ground-glass haze develops later
Granular corneal dystrophy (Type 1 and 2)TGFBI5q31Autosomal Dominant1Childhood; early as 2 years of ageFrequent corneal erosion, photophobia, glare.Decrease in visual acuity as opacification progresses with age. Well-defined granular opacities are observed that don’t extend to the limbus. Type 2 can add snowflakes and lattice lines between the granules.
Macular corneal dystrophy (MCD)CHST616q22Autosomal Recessive1ChildhoodPainful incidence of recurrent corneal erosion, reduced corneal sensitivity, photophobiaSevere visual impairment between 10–30 years.Thinning of the cornea, in advanced stage corneal endothelium is affected and the Descemet membrane develops guttate excrescences. Limbus to limbus stromal haze, which later spreads to superficial, central, elevated white opacities.
Schnyder Corneal Dystrophy (SCD)UBIAD11p36Autosomal Dominant1Childhood to 2nd or 3rd decadeReduced corneal sensitivity, glare increases, disproportionate decrease of photopic vision, may have hyperlipoproteinemia (type IIa, III, or IV)Visual acuity decreases with ageInitial signs include central corneal haze and/or subepithelial crystals (>23 years), arcus lipoids (23–38 years), mid-peripheral panstromal haze (after 38 years)
Congenital Stromal Corneal Dystrophy (CSCD)DCN12q21.33Autosomal Dominant1CongenitalIrregular and cloudy appearance of the cornea, reduced visual acuity, increased glareModerate to severe visual lossDiffuse, bilateral, corneal clouding with flake-like, whitish stromal opacities throughout the stroma, pachymetry demonstrates increase in thickness.
Fleck Corneal Dystrophy (FCD)PIKFYVE2q34Autosomal Dominant1CongenitalAsymptomaticNormalSmall, translucent, discoid opacities that are scattered sparsely throughout without affecting the central cornea. Involvement of the asymmetric or unilateral corneal.
Posterior Amorphous Corneal Dystrophy (PACD)KERA, LUM, DCN, EPYC12q21.33Autosomal Dominant31st Decade, early as 16 weeks, possibly congenital natureMildly effected visual acuityMild visual reductionDiffused grayish-white sheet like opacities in the posterior part of stroma, corneal thinning (>380 µm), flat corneal topography (<41.00 D) and hyperopia in the centroperipheral form.
Descemets Membrane and Endothelial DystrophiesFuchs Endothelial Corneal Dystrophy (FECD); early and late-onsetCOL8A21p34.3–p32, 13pTel–13q12.13, 15q, 18q21.2-q21.32Autosomal Dominant1, 24th decade or laterEpiphora due to recurrent corneal erosion, photophobia, pain, epithelial/stromal edema. Progressive visual impairmentDiffuse thickening of Descemet membrane with excrescences (guttae). Endothelial cells sparse and atrophic
Posterior Polymorphous Corneal Dystrophy (PPCD); Type 1, 2, 3 and 4OVOL2, COL8A2, ZEB1, GRHL220p11.23, 1p34.3–p32.3, 10p11.2, 8q22.3Autosomal Dominant2, 1ChildhoodStromal clouding, endothelial decomposition that is often asymptomaticRarely extensive and progressive visual impairmentDeep corneal lesions that can be nodular, vesicular or blister-like. Edema of the stromal and epithelial layer, endothelial decomposition
Congenital hereditary endothelial dystrophySLC4A1120p13 Autosomal Recessive1CongenitalStromal clouding, blurred vision, photophobiaBlurring visionCorneal thickening, clouding of the cornea, elevated IOP
X-linked Endothelial Corneal DystrophyUnknownXq25X-chromosomal Dominant2CongenitalBlurring visionBlurred vision in malesCloudy cornea only in males, moon crater–like endothelial changes
TGFBI, Transforming growth factor beta-induced; KRT3, Keratin 3; COL8A2, collagen type VIII alpha 2; ZEB1, two-handed zinc-finger homeodomain transcription factor 8; SLC4A11, solute carrier family 4 member 11; PIP5K3, Phosphatidylinositol-3-phosphate/phosphatidylinositol 5-Kinase type III; DCN; Decorin, UBIAD1; UbiA prenyltransferase domain containing 1, CHST6, Carbohydrate sulfotransferase 6 gene; GSN, Gelsolin; TACSTD2, tumor associated calcium signal transducer, KERA, deletion of keratocan; LUM, lumican; EPYC, epiphycan; PIKFYVE, phosphoinositide kinase, FYVE finger containing; OVOL2, Ovo like zinc finger 2; GRHL2, Grainyhead-like transcription factor 2.
Table 2. Advantages and disadvantages of viral and non-viral vectors.
Table 2. Advantages and disadvantages of viral and non-viral vectors.
VectorsCells 12 01280 i001
AdenoVirus
Cells 12 01280 i002
Lenti Virus
Cells 12 01280 i003
AAV Virus
Cells 12 01280 i004
Non-Viral
Advantages
  • Non-mutagenic
  • Highly efficient for infecting dividing and non-diving corneal cells in vivo, in vitro and ex vivo
  • High titre
  • High success rate
  • Highly efficient for infecting and delivering genes into dividing and non-diving corneal cells
  • Sustained gene expression
  • Broad tropism
  • Widely useful for in vitro gene therapy concept testing and expression profiling
  • Larger transgene carrying capacity
  • Highly efficient for infecting and delivering genes into dividing and non-diving corneal cells in vivo, in vitro, and ex vivo
  • Sustained and stable transgene expression
  • Low safety concern
  • Multiple serotypes
  • Minimal/no immune reaction
  • Highly potent and safe for delivering genes into dividing and non-diving cells
  • Sustained and stable transgene expression
  • Minimal/ no immune reaction
  • Large cargo of therapeutic gene transportation capacity
  • Ability to dictate release of therapeutics
Disadvantages
  • Repeat administration of virus is inefficient
  • High immune reaction
  • Common human virus that reduces potency
  • Short term expression
  • High safety concern
  • Highly immunogenic
  • Random integration
potential
  • Safety concern
  • High titer production requires technical skills
  • Small insert size
  • High titer production requires technical skills
  • Concerns of adeno viral contamination
  • Short-term transgene expression
  • Poor efficiency
  • Low-high immune reaction
Table 3. Corneal in vivo and ex vivo gene therapy using AAV.
Table 3. Corneal in vivo and ex vivo gene therapy using AAV.
Adeno-Associated Virus Mediated Corneal Gene Therapy
VectorDiseaseGenePromoterSerotypeDosageModelSpeciesMode of AdministrationOutcomeReference
rAAVGlaucomaMMP3CMVAAV-2/91 × 1011 vgcIn VivoWT mouseIntracameralEfficient transduction resulted in an increase in aqueous concentration and MMP3 activity, also increased outflow facility and decreased IOP[118]
rAAVMPS1-associated Corneal Blindness IDUA CMVAAV8G91 × 1010 vgcEx VivoHumanIntrastromalEfficient widespread transduction resulted in a >10-fold supraphysiological increase in IDUA activity. No significant apoptosis due to AAV or IDUA Overexpression was observed.[119]
rAAVMPS1-associated Corneal Blindness IDUA CMVAAV8G9Increasing volume (50–80 µL); escalating viral dose ranging from (5 × 1010–8 × 1010) per corneaIn VivoMPS I canine modelIntrastromalResolution of corneal clouding as early as 1st week, followed by sustained corneal transparency until the end of 25 weeks in eyes of MPS I canines with advanced disease whereas, prevention against the development of advanced corneal changes while restoring clarity was observed in MPS I canines with early corneal disease.[120]
rAAVCorneal scarringSmad7CAGAAV575 μL; (2.67 × 1013 μg/mL; n = 6)In Vivo2–3-month-old New Zealand White female rabbitsTopical dropsSingle topical application of AAV vectors in rabbit cornea post-PRK led to a significant decrease in corneal fibrosis and corneal haze. [113]
rAAVCorneal scarringDecorinCAGAAV5100 μL; 6.5 × 1012 μg/mLIn Vivo2–3-month-old New Zealand White female rabbitsTopical dropsThe stromal haze and fibrosis were decreased in the corneas infected with rAAV. Decorin virions. No immunogenic or toxic response was observed. [113]
rAAVcorneal inflammation and prevent corneal graft rejectionHLA-G1JETAAV8G92.4 × 1010 vgc total doseIn VivoNaive Lewis ratsIntrastromalCorneal intrastromal delivery of AAV.HLA-G subsequently reduced corneal inflammation, vascularization and fibrosis post-corneal injury.[121]
rAAVCorneal NeovascularizationAngiostatinCMV-5 µL; 1 × 1010 viral particlesIn Vivo10–12 weeks old, Male Sprague-Dawley ratsSubconjunctivalSubconjunctival delivery of AAV. Angiostatin showed a significant reduction of alkali burn-induced corneal angiogenesis. [122]
rAAVCorneal NeovascularizationFlt-1CMVAAV94 × 1011 particles/mLIn VivoRatsAnterior chamberThe subsequent reduction in the development of corneal neovascularization in the stroma of cauterised rats by 36% in comparison to the control group. [123]
rAAVCorneal NeovascularizationmiR-204CMVrAAVrh.103.6 × 1010 GCs (Intrastromal), 3.6 × 1010 GCs (Subconjunctival)In Vivo6- to 8-week-old female C57BL/6J miceIntrastromal injection; subconjunctival injectionAttenuation of corneal neovascularization was observed in the alkali-burned cornea[124]
rAAVCorneal scarring/ fibrosisId3CAGAAV5100 μL; 6.5 × 1012 μg/mLIn Vivo2–3-month-old New Zealand White female rabbitsTopical dropsAttenuation of corneal fibrosis and restoration of corneal transparency was observed. No cellular toxicity was reported.[125]
Table 4. Corneal in vivo gene therapy using LVs.
Table 4. Corneal in vivo gene therapy using LVs.
Lentivirus Mediated Corneal Gene Therapy
VectorDiseaseGenePromoterDosageModelSpeciesMode of AdministrationOutcomeReference
rLVCorneal NeovascularizationEndostatin/Kringle 5CMV50 µL, approximately 108 virus particles/mLIn VivoNew Zealand White rabbitsSubconjunctivalCorneas transduced with rLV.E-K-5 vector demonstrated inhibition of neovascularization and graft failure, whereas in control animals an early onset and profound neovascularization was observed. 5/6 animals in the control group had graft failure. [59]
rLVCorneal Graft RejectionBcl-xLCMV5.5 × 106 IU/mLIn VivoBALB/c mice were used as recipients, and C57BL/6 mice (MHC and multiple minor H disparate) or BALB/c (syngeneic) corneas were used as donorsTopically transducing the cultured corneas ex vivo before transplantation.Delivery of rLV. Bcl-xL to the corneal endothelium of donor corneas significantly improved and promoted allografts’ survival by preventing the endothelium’s apoptosis. [130]
rLVCorneal Graft RejectionPD-L1Ubi-1Between 3.4 × 107 and 1 × 108 titration units (TU)/mLIn VivoMale Lewis (LEW, RT-1) rats served as recipients of male Dark Agouti (DA, RT-1avl) graftsTopically transducing the cultured corneas ex vivo before transplantation.A subsequent increase in the expression of PD-L1 levels in corneal cells helped prolong allograft survival with minimal proinflammatory cytokine expression. [131]
rLVCorneal FibrosisSmad7CMV1 × 104 IFU/μLIn VivoSprague-Dawley rats, approximately 8 weeks of ageTopical dropsExogenous expression of Smad7 gene expression resulted in reduced activation of the TGFβ/Smad signalling caused by the downregulation of phosphorylation of Smad2. Cell proliferation and fibrotic markers were also inhibited by Smad7. [132]
rLVCorneal FibrosisIL-10SV406.3 × 106 TU/mLIn VivoOvineTopically transducing the cultured corneas ex vivo before transplantation.Prolonged survival of corneal allograft by a median of 7 days in recipient corneas transduced with lentivirus containing the therapeutic gene, compared to the control group. [133]
rLVCorneal Fibrosisp35CMV3 × 105 IU/mLIn Vivo8 to 12 weeks old male C57BL/6 (B6) and (B/C) miceTopically transducing the cultured corneas ex vivo before transplantationExogenous expression of the p35 gene reduced the graft-mediated immune response due to decreased CD4+ T-cells expression in the cornea. [134]
Table 5. Corneal in vivo and ex vivo gene therapy using Ads.
Table 5. Corneal in vivo and ex vivo gene therapy using Ads.
Adenovirus Mediated Corneal Gene Therapy
VectorDiseaseGenePromoterDosageModelSpeciesMode of AdministrationOutcomeReference
AdCorneal scarringBMP-7CAG3 µL, 2 × 107 PFU/mLIn VivoC57BL/6 MiceTopicalOverexpression of BMP-7 subsequently reduced the scarring of the alkali burn corneas post-20 days of infection. [138]
AdCorneal scarringSmad7CAG1 × 107 PFU/μLIn VivoC57BL/6 MiceTopicalExogenous expression of Smad7 in the burned corneal tissue resulted in reduced activation or blocking of the Smad3 signalling and nuclear factor-κB signalling via RelA/p65.[139]
AdCorneal NeovascularizationVasohibin-1MMP-11 × 109 PFU/μLIn Vivo6 to 8 weeks old female BALB/c miceSubconjunctivalSubconjunctival delivery of the therapeutic gene significantly reduced the scarring of the alkali burn corneas[140]
AdCorneal NeovascularizationFlk-1CMV2 µL, 3 × 108 PFU/μLIn Vivo6–8 weeks old Sprague–Dawley ratsAnterior chamberSignificant inhibition of neovascularization was observed in the cauterized rat in comparison to the control group. [141]
AdCorneal FibrosisPPAR-γCAG1 × 107 PFU/μLIn Vivo Not specified, possibly topicalUpon overexpression of the transgene in alkali-burned mouse cornea, subsequent suppression of the pro-fibrotic factors and promotion of epithelial healing was observed. [142]
AdCorneal wound healingc-metCMV1 × 108 PFU/μLEx VivoHuman Topically transducing the cultured corneas ex vivoTransduction of diabetic corneas c-met with c-met restored the HGF signalling, normalized the diabetic marker patterns, and accelerated the wound healing process[143]
AdCorneal Graft Rejection CMV1 × 108 PFU/μLIn VivoLewis rats served as recipients of female rats of Dark AgoutiTopically transducing the cultured corneas ex vivo before transplantation. Successful prevention of allogeneic graft rejection in corneal transplantation[56]
AdCorneal Graft RejectionIL-10CMV6.6 × 102–6.6 × 108 PFU/μLIn VivoSheepTopically transducing the cultured corneas ex vivo before transplantation.Overexpression of the immunomodulatory cytokine IL-10 showed significantly prolonged corneal allograft survival and reduced incidence of graft rejection.[144]
AdCorneal Graft RejectionIL-12 p40CMV0.8–1 × 107 PFU/corneaIn VivoSheepTopically transducing the cultured corneas ex vivo before transplantation.Local intraocular production of p40 IL-12 increased the corneal graft survival in comparison to the control group which got rejected at the median of 20 days. [145]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sarkar, S.; Panikker, P.; D’Souza, S.; Shetty, R.; Mohan, R.R.; Ghosh, A. Corneal Regeneration Using Gene Therapy Approaches. Cells 2023, 12, 1280. https://doi.org/10.3390/cells12091280

AMA Style

Sarkar S, Panikker P, D’Souza S, Shetty R, Mohan RR, Ghosh A. Corneal Regeneration Using Gene Therapy Approaches. Cells. 2023; 12(9):1280. https://doi.org/10.3390/cells12091280

Chicago/Turabian Style

Sarkar, Subhradeep, Priyalakshmi Panikker, Sharon D’Souza, Rohit Shetty, Rajiv R. Mohan, and Arkasubhra Ghosh. 2023. "Corneal Regeneration Using Gene Therapy Approaches" Cells 12, no. 9: 1280. https://doi.org/10.3390/cells12091280

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop